Archives

Endocrinology of Pregnancy

ABSTRACT

A coordinated sequence of events must occur in order to establish and successfully maintain a healthy pregnancy. Synchrony between the development of the early embryo and establishment of a receptive endometrium is necessary to allow implantation and subsequent progression of pregnancy. The endocrinology of human pregnancy involves endocrine and metabolic changes that result from physiological alterations at the boundary between mother and fetus. Known as the feto-placental unit (FPU), this interface is a major site of protein and steroid hormone production and secretion. Many of the endocrine and metabolic changes that occur during pregnancy can be directly attributed to hormonal signals originating from the FPU. The initiation and maintenance of pregnancy depends primarily on the interactions of neuronal and hormonal factors. Proper timing of these neuro-endocrine events within and between the placental, fetal, and maternal compartments is critical in directing fetal growth and development and in coordinating the timing of parturition. Maternal adaptations to hormonal changes that occur during pregnancy directly affect the development of the fetus and placenta. Gestational adaptations that take place in pregnancy include establishment of a receptive endometrium; implantation and the maintenance of early pregnancy; modification of the maternal system in order to provide adequate nutritional support for the developing fetus; and preparation for parturition and subsequent lactation.

INTRODUCTION

A coordinated sequence of events must occur in order to establish and successfully maintain a healthy pregnancy.  The endocrinology of human pregnancy involves endocrine and metabolic changes that result from physiological alterations at the boundary between mother and fetus. Known as the feto-placental unit (FPU), this interface is a major site of protein and steroid hormone production and secretion (Figure 1). Additionally, it serves as an endocrine, respiratory, alimentary, and excretory organ, facilitating the exchange of nutrients and metabolic products between the mother and fetus. The fetus is dependent on this effective exchange with the mother for its proper intrauterine growth and development. Thus, it is not surprising that the fetus initiates and influences maternal adaptations to optimize this exchange via complex hormonal mechanisms. Many of the endocrine and metabolic changes that occur during pregnancy can be directly attributed to hormonal signals originating from the FPU. The initiation and maintenance of pregnancy depends primarily on the interactions of neuronal and hormonal factors. Proper timing of these neuro-endocrine events within and between the placental, fetal, and maternal compartments is critical in directing fetal growth and development and in coordinating the timing of parturition. Maternal adaptations to hormonal changes that occur during pregnancy directly affect the development of the fetus and placenta. Gestational adaptations that take place in pregnancy include establishment of a receptive endometrium; implantation and the maintenance of early pregnancy; modification of the maternal system in order to provide adequate nutritional support for the developing fetus; and preparation for parturition and subsequent lactation.

 Figure 1. The interface between mother and fetus, known as the feto-placental unit (FPU), is a major site of protein and steroid hormone production and secretion. 

ENDOMETRIAL RECEPTIVITY 

The menstrual cycle, involves a synchronous production of ovarian steroid hormones, estrogen and progesterone, which induces structural and functional changes within the endometrium in anticipation for embryo implantation and the establishment of a pregnancy.  During the luteal phase, under the primary influence of progesterone, the proliferative endometrium changes into secretory endometrium, which is well vascularized and composed of spiral arteries. A favorable environment for implantation is established via chemokines, growth factors, and cell adhesion molecules (CAMs) produced by the glandular secretory endometrium (1). The chemokines and CAMs serve to attract the blastocyst to the specific sites of implantation where the endometrium is strategically prepared for invasion and placentation (1). When implantation does not occur, a timely regression and destruction of the fully developed endometrium leads to menstruation. However, if implantation occurs, the endometrium continues to grow and undergoes further morphological and molecular changes to provide supportive environment for the growing embryo (2).

Endometrial “receptivity” refers to this physiological state when the endometrium allows a blastocyst to attach, firmly adhere, penetrate, and induce localized changes in the endometrial stroma resulting in decidualization (2). The specific period, known as the “implantation window” opens 4-5 days after endogenous or exogenous progesterone stimulation and closes approximately 9-10 days later (3, 4). Implantation has three stages: apposition, adhesion and penetration. Apposition is an initial unstable adhesion of the blastocyst to the endometrial surface.  This stage is characterized histologically by the appearance of microprotrusions from the apical surface of the epithelium, termed pinopodes, occurring six days after ovulation and retained for 24 hours during the implantation window. The pinopods express chemokines and CAMs, which attract the blastocyst floating within the endometrial cavity to appose.  Additionally, the smooth surface of the pinopodes facilitates the apposition of the blastocyst to the endometrium.  Further encouraging the blastocyst to appose to the pinopods is the removal of adhesion inhibiting mucin, while the areas between pinopods have been shown to express MUC-1, which prevents embryo adhesion (5).  Once the blastocyst is apposed, a stronger attachment is achieved through local paracrine signaling between the embryo and the endometrium. At this stage, the blastocyst is sufficiently adherent to the endometrium as to resist dislocation of the blastocyst by flushing the uterine lumen. The first sign of the attachment reaction coincides with a localized increase in stromal vascular permeability which is manifested as stromal edema at the site of blastocyst attachment (6).  Thus, vascular changes also appear to be an important factor in establishing endometrial receptivity. Following adhesion, the embryo invades through the luminal epithelium into the stroma to establish a relationship with the maternal vasculature. In response to this invasion and the presence of progesterone stimulation, the endometrial stromal cells undergo a process termed decidualization by which they differentiate and become specialized decidual stromal cells. Decidualization is essential for the survival and continued development of the pregnancy. In humans, decidual changes occur throughout the entire endometrium during the luteal phase even in the absence of an embryo, but become widespread in early gestation. These decidual stromal cells are very metabolically active and support the implanting embryo by secreting a wide array of hormones and growth factors including prolactin, relaxin, insulin-like growth factors (IGFs) and insulin growth factor binding proteins (IGFBPs). The endometrial stromal cells are the precursors of decidual stromal cells and appear to originate from both resident uterine mesenchymal stem cells as well as adult bone marrow-derived stem cells (7, 8). Interestingly, bone marrow-derived progenitors have been shown to give rise to functional prolactin-producing decidual stromal cells in decidua of pregnant mice, and appear to play an important role in implantation and pregnancy maintenance (9). The bone marrow is also the source of many leukocytes that infiltrate the endometrium during the secretory phase. In humans, a large influx of leukocytes to the uterus occurs in response to ovulation and rising ovarian P4 production, elevating them to 40% of all endometrial cells in the mid-late secretory phase of the menstrual cycle (10). This gain in leukocyte numbers is primarily due to the accumulation of uterine natural killer (uNK) cells. Studies in mice additionally show that the selected entry of uNK cells into early decidua optimizes angiogenesis and promotes decidual spiral artery vascular remodeling. This influences the timing of uterine lumen closure and thereby the appropriate rate of early fetal development including initiation of trophoblast invasion (11). Macrophages are the second most abundant leukocyte population in the luteal phase endometrium. In addition to uterine NK cells and macrophages, the endometrium contains T cells with no apparent cyclic changes, and rare populations of dendritic cells in luteal phase endometrium, both of which become more abundant in the pregnant decidua. The composition and function of these immune cells at the implantation site and the maternal-fetal interface are highly specialized to foster embryo and placental development and to minimize the chance of immune rejection (12).   

Progesterone is essential in mediating the changes that the endometrium undergoes in the luteal phase in preparation for embryo implantation (13). The effects of progesterone on the uterus have been elucidated through elegant experiments in knockout mice as well as studies using progesterone receptor (PR) antagonists. Mice with global PR knockout are infertile due to defects in ovulation and implantation (14). Their endometrium displays hypertrophy and inflammation of the glandular epithelium associated with failure to undergo decidualization. Mice with a specific knockout in PR-B isoform, however, have normal ovarian function, implantation and reproductive capacity (15, 16). In contrast, mice with a specific knockout in PR-A exhibit lack of decidualization in the endometrial stroma along with endometrial epithelial hyperplasia and inflammation (15, 16), indicating that PR-A is critical for embryo implantation and the normal function of the endometrial epithelium and stroma, while PR-B promotes epithelial hyperplasia of the endometrium. Moreover, administration of the progesterone antagonist mifepristone (RU486) in humans during pregnancy induces abortion, fetal loss or parturition, depending on the gestational age (17, 18). If administered at low doses at the mid- or late follicular phase, it prevents pregnancy by delaying endometrial maturation, while at high doses it delays the LH surge and inhibits ovulation (19, 20).   

The key to endometrial receptivity is the dynamic and precisely controlled molecular and cellular events that involve coordinated effects of autocrine, paracrine, and endocrine factors. Analysis of the transcriptosome of the endometrium during the implantation window using microarray technology has revealed numerous genes that are up- and down-regulated during the “window of implantation” when compared with late proliferative phase endometrium (4, 21). In particular, transcription factors such as the homeobox (HOX) genes are essential for endometrial receptivity by mediating some functions of the sex steroids. HOXA10- and HOXA11-deficient mice have uterine factor infertility due to an implantation defect (22, 23). Both HOXA10 and HOXA11 mRNAs are expressed in human endometrial epithelial and stromal cells; their expression is upregulated by estrogen and progesterone, and is significantly higher in the mid- and late-secretory phases, coinciding with time of embryo implantation (24, 25). As transcription factors, HOX genes regulate other downstream target genes specific to the implantation window, including pinopodes, β3 integrin and insulin-like growth factor-binding protein-1 (IGFBP-1), leading to the proper development of the endometrium and receptivity to implantation (26). Other growth factors, cytokines, and transcription factors produced by the endometrium also assist in the establishment of endometrial receptivity (26, 27).  Impaired endometrial receptivity is considered to be a major limiting factor for the establishment of a pregnancy. Implantation during this time of uterine receptivity is associated with high (85%) success rate for continuing a pregnancy, whereas implantation after cycle day 25 has a much lower success rate (11%) (28).

IMPLANTATION

Pregnancy-related proteins can be found in maternal circulation shortly after fertilization. For example, platelet activating factor (PAF)-like substance, which is produced by the fertilized ovum, is present almost immediately (29-32). After ovulation and fertilization, the embryo remains in the ampullary portion of the fallopian tube for up to 3 days. The embryo undergoes a sequence of cell divisions and differentiation that is not dependent on the hormonal milieu of the fallopian tube or the uterus, as fertilization and early embryonic development occur successfully in vitro. The developing conceptus travels toward the uterus, through the isthmic portion of the tube, for approximately 10 hours, and then enters the uterus as an embryo at the 2- to 8-cell stage (33, 34). With further development, between 3-6 days after fertilization, the embryo becomes a blastocyst floating unattached in the endometrial cavity (34). A schematic representation of the pre-implantation phase of pregnancy is shown in Figure 2. Before implantation, the blastocyst also secretes specific substances that enhance endometrial receptivity. Successful implantation requires precise synchronization between blastocyst development and endometrial maturation. Indeed, there appears to be a cross-talk between the embryo and the endometrium with the endometrium acting as a biosensor that is able to respond favorably to competent embryos but less favorably to incompetent poorly viable embryos destined to fail (35). Ultimately, implantation failure is the result of impaired embryo developmental competence or impaired endometrial receptivity, both having negative effects on the embryo-endometrium cross-talk. It is estimated that embryos account for one third of implantation failures, while suboptimal endometrial receptivity and aberrant embryo-endometrial cross-talk are responsible for the remaining two-thirds (36).

Figure 2. A diagrammatic summary of the ovarian cycle leading to embryo development as it occurs during the first week after fertilization. (Adapted from (37), with permission)

To date, little information exists regarding regulation of steroid production in the embryo. The early embryo and its surrounding cumulus cells secrete detectable estradiol and progesterone well before the time of implantation (38, 39). Mechanical removal of these cells results in the cessation of steroid secretion, while return of the removed cells through co-culture results in restoration of steroid secretion (38). Given this finding, steroid production by the conceptus is thought to be negligible by the time it has reached the endometrial cavity, since it is gradually denuded of cumulus cells as it travels through the fallopian tube.

Conceptus-secreted progesterone may itself affect tubal motility as the conceptus is carried to the uterus (40). Progesterone, by action mediated through catecholamines and prostaglandins (PG), is believed to relax utero-tubal musculature. Moreover, progesterone is thought to be important in tubal-uterine transport of the embryo to the uterine cavity, since receptors for progesterone are found in highest concentrations in the mucosa of the distal one third of the fallopian tube. Estradiol, also secreted by these structures, may balance the progesterone effect so as to maintain the desired level of tubal motility and tone (40). Progesterone antagonizes estrogen-augmented uterine blood flow through depletion of estrogen receptors in the cytoplasm (41). Likewise, estrogen and progesterone also appear to balance one another in the maintenance of blood flow at the implantation site. Both estrogen and progesterone are known to upregulate the expression of multiple angiogenic factors in the uterus, including VEGF, bFGF, PDGF, and TGF-β (42). It is well known that estrogen stimulates an increase in uterine angiogenesis, blood flow and vasodilation by acting both directly on endothelial cells, and/or indirectly on other endometrial cell types via numerous potential promoters (43). In pregnant baboons and sheep, estrogen stimulated uterine and placental blood flow (44). Estrogen treatment significantly increased the paracellular cleft width between endometrial endothelial cells within 6 h considered to result in the increased vascular permeability associated with estrogen administration (45). Unlike estrogen, the angiogenic effects of progesterone in the uterus are believed to occur without concurrent vasodilation (46), as there was no change in endometrial endothelial paracellular cleft width 6 h after progesterone treatment in baboons (45). However, much is still unknown regarding uterine blood flow regulation in pregnancy and how the implanting embryo may influence this process. Human chorionic gonadotropin (hCG) messenger ribonucleic acid (mRNA) is detectable in the blastomeres of 6- to 8-cell embryos; however, it is not detectable in blastocyst culture media until the 6th day (47-49). After implantation is initiated, the embryo is actively secreting hCG, which can be detected in maternal serum as early as the 8th day after ovulation. However, due to the absence of direct vascular communication, secretion of hCG into the maternal circulation is initially limited (50). The primary role of hCG is to prolong the biosynthetic activity of the corpus luteum, which allows continued progesterone production and maintenance of the gestational endometrium. As implantation progresses, the conceptus continues to secrete hCG and other pregnancy-related proteins, and resumes detectable steroid production (38, 39, 51).

Termed trophectoderm (aka outer cell mass), blastomeres lining the periphery of the blastocyst are destined to form the placenta and can be identified at 5 days post-fertilization. The main structural and functional units of the placenta are the chorionic villi, which increase significantly in number during the first trimester of pregnancy. The structure of the chorionic villi is pictured in Figure 3. The villous structure provides a tremendous absorptive surface area to facilitate exchange between the maternal and fetal circulation. The maternal blood arrives from the spiral arteries and circulates through the intervillous space. Fetal blood moves in the core of the chorionic villi within the villous vessels; thus, fetal and maternal blood is never mixed in this system. The key cells inside the chorionic villi are the cytotrophoblasts. They have the ability to proliferate, invade and migrate or to differentiate, through aggregation and fusion, to form a syncytial layer of multi-nucleate cells lining the placental villi, known as the syncytiotrophoblasts.

By 10 days post-fertilization, 2 distinct layers of invading trophoblasts have formed. The inner layer, the cytotrophoblasts, is composed of individual, well-defined and rapidly dividing cells. The outer layer, the syncytiotrophoblasts, is a thicker layer comprised of a continuous cell mass lacking distinct cell borders. Syncytiotrophoblasts line the fetal side of the intervillous space opposite the decidualized endometrium of the maternal side. Immunohistochemically, cytotrophoblasts stain for hypothalamic-like protein hormones: gonadotropin releasing hormone (GnRH), corticotrophin releasing hormone (CRH), and thyrotropin releasing hormone (TRH) (52-64). Juxtaposed syncytotrophoblasts stain immunohistochemically for the corresponding pituitary-like peptide hormones: human chorionic gonadotropin (hCG; analogous to pituitary luteinizing hormone, LH), adrenocorticotropic hormone (ACTH) and human chorionic thyrotropin (hCT). Anatomically, this arrangement suggests that these 2 layers mirror the paracrine relationship of the hypothalamic-pituitary axis (52-64).

Syncytiotrophoblasts, the principal site of placental steroid and protein hormone biosynthesis, have a large surface area and line the intervillous space which exposes them directly to maternal bloodstream without the vascular endothelium and basement membrane which separates them from the fetal circulation (Figure 3-5). This anatomic arrangement explains why placental proteins are secreted almost exclusively into the maternal circulation in concentrations much higher than those in the fetus (65). The syncytiotrophoblast layer contains the abundant subcellular machinery characteristic of cells primarily responsible for hormone synthesis. Amino acids of maternal origin are assembled into pro-hormones. Pro-hormones are then packaged into early secretory granules and transferred across the trophoblastic cell membranes as mature granules. Mature granules become soluble as circulating hormones in maternal blood as they pass through the intervillous space (65).

Figure 3. A. A depiction of a blastocyst implanting in the uterus. B. A longitudinal section of a chorionic villus at the feto-maternal interface at about 10 weeks' gestation. The villous serves as a bridge between maternal and fetal compartments. C. Human placental ultra-structure seen in cross section. Syncytiotrophoblasts line the fetal surface of the intervillous space and interact with the maternal blood supply to secrete placental hormones directly into the circulation. Decidua lines the maternal surface of the intervillous space and secretes protein hormones. (From (66), with permission)

DECIDUA AND DECIDUAL HORMONES

The decidua is the endometrium of pregnancy. Decidualized endometrium is a site of maternal steroid and protein biosynthesis that relates directly to the maintenance and protection of the pregnancy from immunologic rejection. For instance, decidual tissue secretes cortisol, and in combination with hCG and progesterone secreted by the conceptus, cortisol produced by the decidua acts to suppress the maternal immune response conferring the immunologic privilege required by the implanting conceptus (67, 68).

Decidual Prolactin

Decidual prolactin is a peptide hormone having chemical and biological properties identical to pituitary prolactin (69). Prolactin, derived from decidualized endometrium, is first detectable in the endometrium at a time corresponding to implantation-cycle day 23. Progesterone is known to induce decidual prolactin secretion (70). Scant decidual prolactin enters the fetal or maternal circulation after it is transported across the fetal membranes from the adherent decidua and is released into the amniotic fluid (71). Unaffected by bromocriptine administration, decidual production of prolactin takes place independent of dopaminergic control (69).

Decidual prolactin secretion rises in parallel with the gradual rise in maternal serum prolactin seen until 10 weeks’ gestation, then it rises rapidly until 20 weeks, and falls as term approaches (72). Decidua-derived prolactin serves to regulate fluid and electrolyte flux through fetal membranes by reducing permeability of the amnion in the fetal-to-maternal direction (69-71, 73-77). Circulating prolactin in the fetus is secreted by the fetal pituitary gland, while prolactin found in the maternal circulation is secreted by the maternal pituitary gland under the influence of estrogens.  Unlike decidual prolactin, these circulating levels are both suppressed by maternal ingestion of bromocriptine.

Decidual Insulin-like Growth Factor Binding Protein-1 (IGFBP-1)

IGF binding protein-1 (IGFBP-1) is a peptide hormone that originates from decidual stromal cells. In non-pregnant women, circulating IGFBP-1 does not change during cycling of the endometrium, while IGFBP-3 is the main circulating IGFBP. During pregnancy, however, there is a several-fold increase in serum IGFBP-1 levels that begins during the first trimester, peaks during the second trimester, and falls briefly only to peak a second time before term (78). IGFBP-1 inhibits the binding of insulin-like growth factor (IGF) to receptors in the decidua and inhibits fetal growth. Newborn birth weight correlates directly with maternal IGF-1 levels, and inversely with circulating IGFBP-1 levels (79).  

Progesterone-Associated Endometrial Protein (PAEP)

Previously known as pregnancy protein-14, PAEP is a glycoprotein hormone synthesized by secretory and decidualized endometrium that is detectable around cycle day 24 (80). In serum, it rises sharply around cycle day 22 to 24, reaching its peak value at the onset of menstruation; if pregnancy occurs, levels remain high (81). In pregnancy, PAEP rises in parallel with hCG (78). Like hCG, PAEP is thought to have immunosuppressant properties in pregnancy (80). PAEP levels are often low in those patients with conditions, such as ectopic pregnancy, in which there is little decidual tissue produced (82).

PROLONGATION OF CORPUS LUTEUM FUNCTION

Primary steroid products of the corpus luteum are progesterone, 17β-progesterone, estradiol and androstenedione. Low-density lipoprotein (LDL) cholesterol is the main precursor responsible for corpus luteum progesterone production (83). Between 6- and 7-weeks’ gestation, corpus luteum function naturally begins to decline. During this luteal-placental transition period, production of progesterone shifts to the developing placenta (Figure 4).

Pulsatile pituitary LH secretion in the early luteal phase followed by hCG secreted from the implanting conceptus act to stimulate progesterone production from the corpus luteum. Removal of the corpus luteum before 6 weeks of gestation increases the risk of abortion (67a). Thus, regarding early pregnancy, progesterone is considered the most important steroid product in this group because progesterone alone can maintain a pregnancy that would otherwise abort in a lutectomized woman (84). For example, exogenous progesterone, given to an agonadal woman pregnant through egg-donor in vitro fertilization (IVF), maintains the pregnancy through the first trimester until placental progesterone secretion is established (85). For this reason, in patients with corpus luteum dysfunction or in whom the corpus luteum has been removed surgically, supplementation with exogenous progesterone is frequently initiated and extended beyond approximately 10 weeks of gestation, the critical period of the luteal-placental shift.

Figure 4. A shift in progesterone production from the corpus luteum to the placenta occurs at approximately the 7th to 9th week of gestation. The small, shaded area represents the estimated duration of this functional transition. (From (86), with permission)

In women with first-trimester threatened abortion, progesterone concentrations at the time of initial evaluation are often predictive of ultimate outcome (87). Abortion will occur in approximately 80% of those with progesterone concentrations under 10 ng/mL; viable pregnancies are virtually never observed at concentrations of <5.0 ng/mL (88).

Corpus Luteum Relaxin

Relaxin is a peptide hormone produced by the corpus luteum, and not detected in non-pregnant women or men.  Although it is argued that the corpus luteum is the sole source of relaxin in pregnancy, it has also been identified in the placenta, decidua and chorion (89-91). The maternal serum concentrations of relaxin rise during the first trimester, when the corpus luteum is dominant, and decline in the second trimester. Interestingly, when women with a normal pregnancy were compared with pregnant women using egg donor (therefore, no corpus luteum), relaxin was only identified in the women with a pregnancy derived from her own eggs.  However, the duration of pregnancy and labor outcomes were not different between the two groups (92).  The presence of relaxin suggests that it may play a role in early pregnancy, but its function is still unclear. 

In animals, relaxin softens (ripens) the cervix, inhibits uterine contractions, and relaxes the pubic symphysis (93).  These changes are similar to those seen during human labor.  Additionally, in vitro studies of human cervical stromal cells have shown that relaxin induces changes consistent with cervical ripening (94, 95).  Human relaxin primarily binds to relaxin receptors in the decidua and chorionic cytotrophoblasts (96).  Relaxin, originating in the decidua and binding to its receptors in the fetal membranes, increases cytokine levels that can activate matrix metalloproteinases and lead to rupture of fetal membranes and labor (97). Thus, relaxin may play a facilitatory role in labor, however its role is still not clearly defined.  

PLACENTAL COMPARTMENT

Unique to mammals, the placenta plays a major role in balancing fetal growth and development with maternal homeostasis. The fetus develops in an environment where respiration, alimentation and excretory functions are provided by the placenta. The human placenta is hemochorial, which means the chorion is in direct contact with maternal blood. Cyto- and syncytiotrophoblast cells of the placenta have direct access to the maternal circulation.  In contrast, the trophoblast layer prevents most maternal hormones from entering the fetal compartment, and consequently the fetal/placental endocrine system generally develops and functions independently of that of the mother.  Over time, the placenta has evolved as a system through which viviparity or livebirth could take place with dependable success.

The placenta functions, to some extent, as a hypothalamic-pituitary-end organ-like entity owing to the inherent ability of this type of system, with its stimulatory and inhibitory feedback mechanisms, to dynamically regulate factors that affect fetal growth and development under a variety of conditions. In the fully developed hypothalamic-pituitary-end organ schema of humans, neural inputs to the hypothalamus serve to regulate the secretion of hypothalamic releasing hormone peptides. However, in the placenta there are no such direct neural inputs, and the exact mechanism(s) responsible for regulation of the secretion of hypothalamic-like placental peptides is unknown.

Changes in maternal hormone concentrations play a critical role in modulating the metabolic and immunologic changes required for successful outcome in pregnancy. The fetus and placenta produce and secrete steroids and peptides into the maternal circulation as well as stimulate maternal hormone production. The origins and amounts of the fetal and placental hormones secreted during pregnancy changes dramatically over the course of the gestational period. Some of the pregnancy-related protein hormones previously discussed are, in part, responsible for the altered steroid concentrations typical of pregnancy.

Placental Steroid Hormones

The placenta is a site of active steroidogenesis which depends on highly integrated and active interactions with both mother and fetus. This is consequent to an elegant complementary of enzymatic deficiencies between placental and fetal compartments (Table 1). The placenta is characterized by significant aromatase, sulfatase, and 11b-hydroxysteroid dehydrogenase type 2 activities juxtaposed with a lack of P450C17 (17a-hydroxylase and 17/20 lyase) activity.

Table 1. Enzymatic Limitations by Compartment

Fetal

Placental

3b-hydroxysteroid dehydrogenase

17a-hydroxylase

 

StAR protein

17/20 lyase

16α-hydroxylase

PLACENTAL PROGESTERONE

The placenta is the main source of progesterone during pregnancy. From the luteal phase to term, maternal progesterone levels rise six- to eight-fold. (Figures 5 and 8) Although, progesterone originates almost entirely from the corpus luteum before 6 weeks' gestational age, its production shifts more to the placenta after the 7th week. Beyond 10 weeks, the placenta is the major definitive source of progesterone (51, 98).

While the placenta produces large amounts of progesterone, it has a limited capacity to synthesize cholesterol de novo (Figure 7). Maternal cholesterol enters the trophoblasts in the form of low-density lipoprotein (LDL) cholesterol which serves as the principal precursor for the biosynthesis of progesterone by the placenta (51, 83, 99). The fetal contribution of progesterone is negligible. This is evident as progesterone levels remain high even after fetal demise.  In the non-human primate estrogen regulates placental progesterone production (100). Progesterone concentrations are less than 1 ng/mL during the follicular phase of the normal menstrual cycle (101, 102). However, in the luteal phase of cycles in which fertilization occurs, progesterone concentrations rise from about 1-2 ng/mL on the day of the LH surge to a plateau of approximately 10-35 ng/mL over the subsequent 7 days. Concentrations remain within this luteal-phase range from the 10th week from the last menstrual flow, and then show a sustained rise that continues until term (Figure 5). At term, progesterone concentrations can range from 100-300 ng/mL (51). Most of the progesterone produced in the placenta enters the maternal circulation.

Figure 5. Relative values of circulating concentrations (mean ±SEM) of progesterone and 17α-progesterone during the course of human pregnancy from fertilization to term. The data displayed demonstrates values before and after the luteinizing hormone (LH) surge. Gestational ages are calculated from last menstrual flow. (Adapted from (103), with permission)

The human deciduas and fetal membranes also synthesize and metabolize progesterone (104).  In this case, neither cholesterol nor LDL-cholesterol are significant substrates; pregnenolone sulfate may be the most important precursor.  Progesterone has been shown to exert important functions in implantation and parturition to include promotion of endometrial decidualization; inhibition of smooth muscle contractility; decrease in prostaglandin (PG) formation, which helps maintain myometrial quiescence and prevent the onset of uterine contractions; and inhibition of immune responses like those involved in graft rejection. It is believed to work in concert with hCG and decidual cortisol to inhibit T-lymphocyte-mediated tissue rejection and confer immunologic privilege to the implanted conceptus and developing placenta (105, 106). In animal models, progesterone extends the survival of transplanted human trophoblasts, and high intervillous concentrations of progesterone are of major importance in blocking the cellular immune rejection of the foreign antigens originating from the pregnancy (106).

In addition to its roles in endometrial and myometrial function, progesterone also serves as a substrate for fetal adrenal gland production of glucocorticoids (cortisol) and mineralocorticoids (aldosterone) (107). This important function is consequent to the deficiency of 3b-hydroxysteroid dehydrogenase (3b-HSD) activity in the active fetal zone of the fetal adrenal gland.

PLACENTAL 17α-HYDROXYPROGESTERONE

Like progesterone, during the first several weeks of gestation and through the time of the luteal-placental shift, 17α-hydroxyprogesterone concentrations primarily reflect the steroidogenic status of the corpus luteum (108). However, by the tenth week of gestation, 17 α-hydroxyprogesterone has returned to baseline levels, indicating that the placenta has little 17 α-hydroxylase activity.  During the third trimester the placenta uses fetal D5-sulfoconjugated precursors to secrete increasing amounts of 17α-hydroxyprogesterone, and at this point the placenta becomes the major source of this hormone at term (108).

Concentrations of 17α-hydroxyprogesterone are less than 0.5 ng/mL during the follicular phase of normal menstrual cycles. In cycles leading to pregnancy, 17α-hydroxyprogesterone concentrations rise to about 1 ng/ml on the day of the LH surge, decline slightly for about 1 day, and rise again over the subsequent 4-5 days reaching a level of 1-2 ng/ml. Concentrations then increase slightly to a mean of approximately 2 ng/ml (luteal phase levels) by the end of the 12th week. This level remains stable until a gestational age of about 32 weeks at which time it begins an abrupt, sustained rise at about 37 weeks to approximately 7 ng/ml, a level that persists until term (108) (Figures 5 and 8). The rise in 17α-hydroxyprogesterone that begins at 32 weeks strongly correlates with the fetal maturational processes known to begin at this time. Hence, 17α-hydroxyprogesterone concentration exhibits a bimodal pattern in normal pregnancy.

PLACENTAL 17β-ESTRADIOL

The corpus luteum is the exclusive source of 17β-estradiol during the first 5-6 weeks of gestation. After the first trimester, the placenta is the major source of circulating 17β-estradiol (51). The rate of estrogen production and the level of circulating estrogens increase markedly during pregnancy. Concentrations of 17β-estradiol are less than 0.1 ng/mL during the follicular phase of the cycle and reach about 0.4 ng/mL during the luteal phase of normal menstrual cycles (101). Following fertilization, 17β-estradiol increases gradually to a range of 6-30 ng/mL at term (102) (Figures 6 and 8). Because it is deficient in 17-hydroxylase enzyme activity and 17-20 desmolase (lyase) activity, the placenta is unable to convert progestogens to estrogens. Thus, the placenta relies on 19-carbon androgen precursors produced by the fetal and maternal adrenal glands. Sources of estrogen biosynthesis by the maternal-fetal-placental unit are depicted in Figure 8. The major source of fetal adrenal dihydroepiandrostenediene sulfate (DHEAS) is LDL-cholesterol circulating in the fetal blood. A minor source of fetal adrenal DHEAS is derived from pregnenolone secreted by the placenta. Twenty percent of fetal cholesterol is derived from the maternal compartment. Since amniotic fluid cholesterol levels are negligible, the main source of cholesterol is the fetal liver. As gestation advances, increasing quantities of 17β-estradiol are synthesized from the conversion of circulating maternal and fetal DHEAS by the placenta. At term, approximately equal amounts of estrogens are produced from circulating maternal DHEAS and fetal DHEAS (51, 109). The fetal endocrine system is notable for extensive conjugation of steroids with sulfate. Consequently, the placenta relies on sulfatase activity to cleave sulfate conjugates in the fetal compartment. Naturally occurring placental sulfatase deficiency results in a low estrogen state in pregnancy (110).

The cytochrome P450 aromatase enzyme is responsible for converting 19-carbon precursors to estrogen (111). The efficiency of this enzyme affords the fetus protection from virilization even in the presence of large amounts of aromatizable androgens.

Figure 6. Relative values of circulating concentrations (mean ±SEM) of 17β-estradiol, estriol and estrone during the course of human pregnancy from fertilization to term. Data displayed demonstrate values before and after the luteinizing hormone (LH) surge. Gestational ages are calculated from last menstrual flow. (Adapted from (112), with permission)

The vasodilatory function of estrogens in pregnancy are well described. In animal models, direct estrogen injection into the uterine arteries produces striking increases in blood flow. Without question, 17β-estradiol is the most potent estrogen in this role. Estriol and estrone, though less active, also produce this effect (113). Because the exposure of the utero-placental bed to direct estriol secretion is enormous, estriol may be the principal up-regulator of uterine blood flow. This may be the dominant role of estriol in human pregnancy (113). Estrogen regulated mechanisms may also allow the fetus to govern production and secretion of progesterone during the third trimester. In primates, estrogen regulates the biosynthesis of placental progesterone by regulating the availability of LDL-cholesterol for conversion to pregnenolone and its downstream steroid products (114). Estrogens are also thought to contribute to mammary gland development and fetal adrenal gland function.

PLACENTAL ESTRADIOL

Estriol is first detectable in maternal serum at 9 weeks of gestation (51, 109, 115, 116). This temporal relationship closely corresponds to the early stages of steroidogenic maturation in the fetal adrenal cortex (51). Hence, the continued production of estriol is dependent upon the presence of a living fetus. Concentrations of estriol are less than 0.01 ng/ml in non-pregnant women. First detectable at approximately 0.05 ng/ml by 9 weeks, estriol increases gradually to a range of approximately 10-30 ng/ml at term (51, 98, 115, 117). Between 35- and 40-weeks gestational age, estriol concentrations increase sharply in a pattern that reflects a final surge of intrauterine steroidogenesis just prior to term (Figures 6 and 8).

Figure 7. Synthesis of estrogen and progesterone within and between the maternal, placental and fetal compartments. (Adapted from (118), with permission)

The placenta lacks 16a-hydroxylase activity and consequently, estriol with its 16a-hydroxyl group, must be synthesized from an immediate fetal precursor. The fetal liver provides 16a-hydroxylation of DHEAS for placental estriol synthesis. Interestingly, hepatic 16a-hydroxylation activity disappears postnatally.

Figure 8. Circulating maternal steroid hormone levels throughout early pregnancy. The first-trimester relationship of these steroid hormones to human chorionic gonadotropin (hCG) is shown.

Progestogens
Progesterone o--o--o-
17-a-hydroxyprogesterone -Δ-Δ-Δ-

Estrogens
17-
β-estradiol ---
Estriol -o-o-o-
Estrone -x-x-x-

Human chorionic gonadotropin (hCG)
-Δ-Δ-Δ- 
(From ref. 89, with permission)

PLACENTAL ESTRONE

For the first 4-6 weeks of pregnancy, estrone originates primarily from maternal sources such as the ovaries, adrenals, or peripheral conversion (102). Later, the placenta secretes increasing quantities of estrone from the conversion of circulating maternal and fetal DHEAS. The placenta continues to be the major source of circulating estrone for the remainder of the pregnancy (51). Estrone concentrations are less than 0.1 ng/mL during the follicular phase and may reach a maximum of 0.3 ng/mL during the luteal phase of a normal menstrual cycle. Following fertilization, estrone concentrations remain within the luteal phase range through weeks 6-10 of gestation (98). Subsequently there is a gradual increase to a wide range of 2-30 ng/ml at term (51, 98, 102) (Figures 6 and 8). In the absence of fetal adrenal gland function, as in the case of anencephaly, maternal estrogen levels are extremely low, suggesting that the maternal contribution of DHEAS to total estrogen synthesis is negligible.

Placental Protein Hormones

As detailed previously, the placental cytotrophoblast-syncytiotrophoblast relationship mirrors the hypothalamic-pituitary system. The surface of the syncytiotrophoblast is in direct contact with maternal blood within the intervillous space, and consequently, placental proteins are preferentially secreted into the maternal compartment. Table 2 outlines the various peptides associated with the endocrinology of human pregnancy.

Table 2. Pregnancy Specific Protein Hormones by Compartment

Fetal

Placental

Maternal

Alpha-fetoprotein

Hypothalamic-like (cytotrophoblast)         

- GnRH                                                

- CRH                                     

- TRH                                     

- GHRH                                  

- Somatostatin           

Pituitary-like (syncytiotrophoblast)           

- hCG

- hGH

- ACTH

- hPL

- hCT                                      

- Oxytocin

Growth factors

- Inhibin

- Activin                                              

- IGF-I/IGF-II

Other proteins

- Pregnancy specific β1-glycoprotein

- PAPP-A

Decidual derived

-Prolactin

-IGFBP-1

-PP14

Corpus luteum derived

-Relaxin

PLACENTAL PROTEINS: HYPOTHALAMIC-LIKE PROTEINS

Placental Gonadotropin Releasing Hormone (GnRH)

Gonadotropin releasing hormone derived from the placenta is biologically and immunologically similar to the hypothalamic decapeptide GnRH (54). Gonadotropin releasing hormone activity has been localized to the cytotrophoblast cells along the outer surface of the syncytiocytotrophoblast layer. Human chorionic gonadotropin (hCG) has been localized to the adjacent syncytiocytotrophoblast layer. GnRH production peaks at about 8 weeks’ gestation and then decreases as the pregnancy advances in gestational age (54-57). Furthermore, GnRH levels parallel those of hCG in both the placenta and maternal circulation (57).

Placental GnRH stimulates hCG release through a dose-dependent, paracrine mechanism (119). There is little augmentation of hCG secretion by GnRH in first trimester placental culture, because hCG production is already close to maximum (57). In contrast, at mid-trimester there is a marked dose-dependent GnRH augmentation of hCG release in vitro, with this effect diminishing in the term placenta. Likely due to the low affinity of placental GnRH receptors and dilution effect of the maternal circulation, intravenous administration of GnRH during pregnancy does not increase serum hCG. Thus, it seems most likely that locally produced placental GnRH is responsible for stimulation of placental hCG production via paracrine mechanisms (119). GnRH release is increased by estrogen, activin-A, insulin and prostaglandins, and inhibited by progesterone, inhibin, follistatin and endogenous opiates (120).

Placental Corticotrophin Releasing Hormone (CRH)


Placental CRH is structurally similar to the hypothalamic peptide, CRH (121, 122). Due to this similarity, it is easily measured in amniotic fluid as well as fetal and maternal plasma. Pro-CRH mRNA is present in cytotrophoblasts (123). CRH is also intensely immunoreactive in the decidua (53). CRH is found in maternal serum at low levels during the first and second trimesters of uncomplicated pregnancies, but rises dramatically in the third trimester of normal gestations or earlier if there are pregnancy complications resulting from such factors as prematurity, diabetes, or hypertension.(124). The highest levels of CRH are found at labor and delivery. Although concentrations of CRH in fetal plasma are lower than those found in maternal plasma, there exists a significant correlation between maternal and fetal plasma CRH (124). There is a 3-fold rise, in amniotic fluid CRH between the second and third trimester (124, 125). Placenta-derived CRH stimulates placental ACTH release in a dose-dependent manner in vitro (126, 127). Corticotrophin releasing hormone and ACTH are both released into fetal and maternal circulation; their activity is moderated by maternal CRH binding proteins (124).

Placental CRH participates in the surge of fetal glucocorticoids associated with late third trimester fetal maturation (124, 126, 128). When uterine blood flow is restricted, secretion of both CRH and ACTH is increased. Corticotrophin releasing hormone is a potent utero-placental vasodilator (129, 130). Corticotrophin releasing hormone is released into the fetal circulation in response to fetal stress and in conditions leading to fetal growth restriction (131-133). High circulating maternal CRH is believed to be responsible for the elevated plasma ACTH and cortisol found in pregnancy, which renders them unresponsive to feedback suppression of plasma cortisol (124-126, 128, 134). Corticotrophin releasing hormone stimulates prostaglandin synthesis in fetal membranes and placenta. In pre-eclampsia, fetal asphyxia, premature labor, and other conditions leading to fetal growth restriction CRH is frequently elevated (131-133).

Placental Thyrotropin Releasing Hormone

Thyrotropin releasing hormone is found in the cytotrophoblast layer; however, this molecule is different from the tripeptide produced by the hypothalamus (135). It is localized primarily in the syncytiotrophoblast but also in the fetal and maternal blood vessels as well as in the extravillous trophoblast. The concentration of TRH is higher in the fetal circulation, which is likely due rapid protease degradation on the maternal side (136). Since hCG is regarded as the principal placenta-derived thyroid stimulator, a significant role for TRH is uncertain, although it may be involved in thyroid function regulation during fetal life (137).

Placental Growth Hormone Releasing Hormone (GHRH)

GHRH has also been identified in the human placenta, but its cellular localization and function are unknown (126). Its structure is identical to that of the hypothalamic GHRH peptide. The levels of placental GHRH do not contribute to maternal circulating levels of the extra villous the presence of GHRH receptor in the placenta GHRH does not regulate placental growth hormone production. 

Somatostatin (SRIF)  

Somatostatin (SRIF) is a peptide that exerts a variety of regulatory actions interacting with G protein-coupled receptors. Placental somatostatin has been found in early pregnancy villi, cytotrophoblast and in the decidua; while its binding sites have been identified in term placental membranes and cytotrophoblast (64, 138, 139). The amount of placental somatostatin decreases with increasing gestation and it does not contribute to maternal circulating levels of the peptide.  The role of placental somatostatin remains unclear.

PLACENTAL PROTEINS: PITUITARY-LIKE HORMONES

Placental Human Chrorionic Gonadotropin (hCG)

Human chorionic gonadotropin is a glycoprotein structurally similar to follicle stimulating hormone (FSH), luteinizing hormone (LH), and thyroid stimulating hormone (TSH). It is similar to luteinizing hormone (LH) in action. As is true of the other glycoprotein hormones, hCG is composed of 2 non-identical subunits that associate non-covalently (52, 140). The α subunit consists of an amino acid sequence essentially identical to and shared with the other pituitary glycoprotein hormones. On the other hand, the β subunit is structurally similar to the α subunit yet it differs enough to confer specific biologic activity on the intact dimeric hormone. The subunits differ primarily at the carboxyl terminus where the β subunit of hCG has a 30-amino-acid tailpiece that is not present in the human LH β subunit. Glycosylation in this region of HCG accounts for the longer half-life (32-37 hours) of hCG relative to LH (24h vs. 2h, respectively). The molecular weight of the hCG dimer is estimated at 36.7 kDa with the α subunit contributing 14.5 kDa and the β subunit 22.2 kDa (140). The hCG α subunit is found in the cytotrophoblast layer only (57, 60).

As mentioned previously, hCG mRNA is detectable in embryos as early as the 6- to 8-cell stage (47). After implantation of the conceptus, hCG is detectable in the syncytiotrophoblast layer (outer trophectoderm layer) (57, 60-62). Human chorionic gonadotropin is secreted by the syncytiotrophoblasts of the placenta into both the fetal and maternal circulation. Plasma levels increase, doubling in concentration every 2-3 days between 60 and 90 days of gestation. At 3-4 weeks' gestation, the mean doubling time of dimeric hCG is 2.0 ±1.0 days and increases to about 3.5 ±1.5 days at 9-10 weeks (57). The average peak hCG level is approximately 110,000 mIU/mL and occurs at 10 weeks’ gestation (57). Between 12 and 16 weeks, average hCG decreases rapidly with the concentration halving every 2.5 ±1.0 days before reaching 25% of first trimester peak values. Levels continue to fall from 16 to 22 weeks at a slower rate (mean halving rate of 4.0 ±2.0 days) to become approximately 10% of peak first trimester values (57). During the third trimester mean hCG levels rise in gradual, yet significant, manner from 22 weeks until term (57). Interestingly, hCG levels are comparatively higher in women bearing female fetuses.

Human chorionic gonadotropin secretion is related directly to the mass of hCG-secreting trophoblastic tissues. In vivo, the release of hCG has been correlated with the widths of trophoblast tissue from 4 to 20 weeks and with placental weight from 20 to 38 weeks, respectively (57). The rapidly rising hCG seen between 3-4 and 9-10 weeks’ gestation coincides with the proliferation of immature trophoblastic villi and the extent of the syncytial layer (57). As expected, declining hCG levels are associated with a relative reduction in the mass of the syncytiotrophoblast and cytotrophoblast tissue. From 20-22 weeks until term a gradual increase in dimeric hCG corresponds with a similar increase in placental weight and villus volume (57).

Thus, in early gestation rising hCG levels reflect the histological finding of a rapidly proliferating and increasingly invasive placenta. Later in pregnancy, declining hCG levels are associated with a relative reduction in the number and mass of trophoblasts; therefore, hCG levels mirror the placenta's morphologic transformation from an organ of invasion to an organ of transfer (57).

Levels of the β subunit of hCG mirror those of dimeric hCG. The α subunit, undetectable until around 6 weeks' gestation, rises in a sigmoid fashion to reach peak levels at 36 weeks. Levels of the individual subunits are very low relative to dimeric hCG; they are approximately 2,000-fold to 150-fold less than dimeric forms at 6 and 35 weeks, respectively) (57).

With respect to the regulation of hCG production and secretion, hCG secretion appears to be related to placental GnRH release (119). In vitro, hCG is released in pulses at a frequency and amplitude that correlate with the release of placental GnRH (119). In addition, hCG production is stimulated by glucocorticoids and suppressed by DHEAS (126). In vitro, cyclic AMP (cAMP) analogues augment hCG secretion. In humans, similar to pituitary secretion of gonadotropins, decidual inhibin and prolactin inhibit hCG production by term trophoblasts whereas decidua-derived activin augments it (140, 141), with stimulation by estrogen and a negative feedback by progesterone.

Human chorionic gonadotropin, the primary luteotropic factor involved in supporting and maintaining the corpus luteum, ensures the continuous secretion of progesterone until the placenta can perform this function (142). It has immunosuppressive properties, likely involving maternal T-lymphocyte function and it possesses thyrotropic activity (143). Human chorionic gonadotropin may stimulate steroidogenesis in the early fetal testes resulting in virilization and sexual differentiation in males (144, 145). The functions of hCG are summarized in Figure 9.

Figure 9. The physiological roles of human chorionic gonadotropin (hCG) during the course of human pregnancy from fertilization to term. (Adapted from (146), with permission)

Placental Growth Hormone (GH)

Growth hormone is a single-chain peptide hormone structurally related to prolactin and human chorionic somatomammotropin (hCS). Up to the first 15-20 weeks of pregnancy, pituitary growth hormone (GH) is the main form present in the maternal circulation. From 15-20 weeks to term, placental GH gradually replaces pituitary GH, which eventually becomes undetectable (147-151). In contrast to the pulsatile output of pituitary GH, the daily profile of placental GH release is non-pulsatile (150). Syncytiotrophoblasts directly bathing in maternal blood are the site of placental GH synthesis. This cell layer is the placental site of the major glucose transporter, Glut1, and responds to rapid variations in maternal blood glucose levels by modifying placental GH secretion (152, 153).

The rate of secretion of pituitary GH is known to change rapidly, depending on the net result of multiple stimulatory and inhibitory input. The regulation of placental GH is quite different. The rate of synthesis of placental GH, and thus the maternal circulating levels, increases with the growth of the placenta (154). Growth hormone releasing hormone (GHRH) does not modulate placental GH expression in vitro, in vivo, or in the presence of glucose (155, 156). Figure 10 shows both the stimulatory and inhibitory mediators of maternal pituitary GH output, including the influence of placental growth hormone.

Production of maternal insulin-like growth factor-1 (IGF-I) is regulated by placental growth hormone. IGF-1 concentrations in the maternal plasma, studied in a large number of pregnancies, correlate with the corresponding placental GH. The IGF-1 levels do not vary significantly during the first weeks of gestation, but then increase gradually from 165 ±44.5 mg/L at about 24-25 weeks' gestation, and reach levels of 330.5 ±63.5 mg/L in a manner similar to the increases seen in placental GH. It should be noted that circulating maternal IGF-I levels also reflect placental IGF-I secretion. This growth factor, however, does not appear to be strongly expressed in human placenta; in particular; it is not expressed in the syncytiotrophoblast cell layer (157).

The biologic activities of GH and related peptide hormones can be classified into two general categories: somatogenic and lactogenic. Somatogenic activities are related to linear bone growth and alterations in carbohydrate metabolism (158, 159). The primary function of GH is to protect nutrient availability for the fetus. Via local and hepatic IGF-1, placental GH stimulates gluconeogenesis and lipolysis in the maternal compartment.

Figure 10. Shown is a representation of the hypothalamic-growth hormone-IGF-I axis, with details of its modification during pregnancy. A. In the non-pregnant state, pituitary GH secretion is regulated through hypothalamic control. Pituitary GH regulates the secretion of IGF-I, which, in turn, exerts negative feedback action on GH at the hypothalamic-pituitary level. B. During the latter half of pregnancy, the GH-IGF axis is inhibited by large amounts of estrogen. The large increase in placental GH exerts an inhibitory effect on GH secretion mediated by placental GH on the hypothalamus and pituitary. (From (160), with permission)

Placental Human Placental Lactogen (hPL), [Human Chorionic Somatomammotropin (hCS)]

Human placental lactogen is a single-chain polypeptide with two intramolecular disulfide bridges. The structures of hPL, prolactin, and growth hormone are very similar. Eighty-five percent of its amino acids are identical to human pituitary growth hormone and human pituitary prolactin (69, 161). Furthermore, hPL shares biologic properties with both growth hormone and prolactin (69, 161). Thus, it has primarily lactogenic activity but also exhibits some growth hormone-like activity; therefore, it is also referred to as chorionic growth hormone (hCGH) or human chorionic somatomammotropin (hCS). Human placental lactogen is secreted from the syncytiotrophoblast cell layer. Unlike hCG concentrations, levels of hPL rise with advancing gestational age and plateau at term. Human placental lactogen is first detectable during the fifth week of gestation, and rises throughout pregnancy maintaining a constant hormone weight to placenta weight relationship (162). Concentrations reach their highest levels during the third trimester, rising from approximately 3.5 µg/mL to 25 µg/mL at term (162). Although the level of hPL in serum at term is the highest of all placenta-derived protein hormones, its clearance from the circulation is so rapid that it cannot be detected after the first post-partum day.

Since hPL is secreted primarily into the maternal circulation, most of its functions occur at sites of action in maternal tissues. Human placental lactogen is thought to be responsible for the marked rise in maternal plasma IGF-1 concentrations as the pregnancy approaches term (162-164). Human placental lactogen exerts metabolic effects during pregnancy via IGF-I. It is associated with insulin resistance, enhances insulin secretion which stimulates lipolysis, increases circulating free fatty acids, and inhibits gluconeogenesis; in effect, it antagonizes insulin action, induces glucose intolerance, as well as lipolysis and proteolysis in the maternal system (69). In response to fasting and glucose loading, hPL levels rise and fall (162). These metabolic effects favor the transport of ketones and glucose to the fetus in the fasting and fed states, respectively.

Circulating levels of glucose and amino acids are reduced, while levels of free fatty acids, ketones, and triglycerides are increased. The secretion of insulin is augmented in response to a glucose load. The fuel requirements of the developing fetus are met primarily by glucose. It provides the energy needed for protein synthesis and serves as a precursor for the fat synthesis and glycogen formation. Fetal blood glucose levels are generally 10-20 mg/100 ml below those of the maternal circulation; thus, diffusion and facilitated transport favor the net movement of glucose from mother to fetus.

Pregnancy is associated with profound alterations in maternal metabolism. The fetal-maternal relationship favors glucose use by the fetus and forces the maternal tissues to increase their use of alternative energy sources. The endocrine hallmark of this hormonal environment is insulin resistance. Several hormones prevalent during pregnancy are believed to responsible for this altered milieu: estrogens, progesterone, glucocorticoids, human placental lactogen (hPL) and placental GH. Additionally, placental cytokines such as tumor necrosis factor-alpha (TNF-α) contribute to this metabolic state (165).

Placental Adrenocorticotropic Hormone (ACTH)

Placental ACTH is structurally similar to pituitary ACTH (166-178). Under the paracrine influence of placental CRH released from proximal cytotrophoblasts, placental ACTH is secreted by syncytiotrophoblasts into the maternal circulation (179-181). Circulating maternal ACTH is increased above non-pregnancy levels, but still remains within the normal range (182, 183).

Placental ACTH stimulates an increase in circulating maternal free cortisol that is resistant to dexamethasone suppression (179, 182). Thus, relative hypercortisolism in pregnancy occurs despite high-normal ACTH concentrations. This situation is possible due to two main differences in endocrine relationships during pregnancy. First, the maternal response to exogenous CRH is blunted (182). Second, a paradoxical relationship exists between placental CRH, ACTH, and their end-organ product, cortisol; glucocorticoids augment placental CRH and ACTH secretion, not suppress it (127, 180). This positive feedback mechanism allows an increase in glucocorticoid secretion in times of stress in excess of the amount necessary if the mother were not pregnant (127).

Placental Human Chorionic Thyrotropin (hCT)

Human chorionic thyrotropin is structurally similar to pituitary TSH, but it does not possess the common α subunit (135). The placental content of hCT is very small (58). Human chorionic gonadotropin possesses 1/4000th of the thyrotropic activity of TSH, and is thought to exert a more significant effect on the maternal thyroid than does hCT (137), particularly in conditions with high hCG levels such as trophoblastic disease.

PLACENTAL PROTEINS: GROWTH FACTORS  

Placental Inhibin/Activin/Follistatin  

Inhibin and activin are heterodimeric glycoproteins with the former comprised of an α and β subunit and the latter composed of two β subunits.  Inhibin is secreted by the corpus luteum and is present in decidualized endometrium (184, 185). Inhibin and activin dimers have been localized to the syncytiotrophoblast layer, but their individual subunits have been localized to both cytotrophoblasts and syncytiotrophoblasts (186).

Inhibin begins to increase in the maternal circulation above non-pregnant levels by 12 days post-fertilization, dramatically increasing at about 5 weeks' gestation to peak at 8-10 weeks. Subsequently, levels decrease at 12-13 weeks and stabilize until around 30 weeks before they rise again as term approaches (185). The early fluctuations in inhibin levels reflect release from the corpus luteum, whereas the increase seen in the third trimester originates from the placenta and decidua. After delivery, inhibin is undetectable. The inhibin A dimer is the principal bioactive inhibin secreted during pregnancy. Quantification of inhibin A is part of the prenatal quad screen that can be administered during pregnancy at a gestational age of 16–18 weeks. An elevated inhibin A (along with an increased beta-hCG, decreased AFP, and a decreased estriol) is suggestive of the presence of a fetus with Down syndrome.

Activin-A is the major trophoblastic activin product, which similarly increases in maternal circulation throughout pregnancy and peaks at term (187). Interestingly, higher levels of activin-A are found in mid-gestation in women with preeclampsia (188, 189).  Similar to their action in the ovarian follicle, inhibin and activin are regulators within the placenta for the production of GnRH, HCG, and steroids; as expected, activin is stimulatory and inhibin is inhibitory. 

Follistatin is the activin-binding protein expressed in placenta, membranes, and decidua (190). Since follistatin binds activin, it antagonizes the stimulatory effects of activin on placental steroid and peptide production. 

Placental Insulin-Like Growth Factors-I and II (IGF I and II)

Without question, the most important site of IGF-I and IGF-II production is the placenta (191). IGF-I and IGF-II are involved in prenatal growth and development. These growth factors do not cross the placenta into the fetal circulation; however, they may be involved in placental growth (192, 193).  An increase in maternal IGF-I levels during pregnancy with a rapid decrease after delivery indicates a significant placental influence.  There is however, no change in IGF-II levels throughout pregnancy.  In animal studies, the IGF-I produced in the placenta regulates the transfer of nutrients across the placenta to the fetus and thus enhances growth.  Interestingly, neonates with intrauterine growth restriction have reduced levels of IGF-I. IGF-II secreted by the placenta is also important in influencing β cell sensitivity to glucose and modulation of maternal insulin and glucose concentrations during pregnancy (194).

Placental Soluble FMS-Like Tyrosine Kinase (SFLT-1) and Souble Endoglin (sENG)

Soluble Flt-1 is a circulating splice variant of Flt-1, the receptor for VEGF and placental growth factor (PLGF), while sENG is the circulating receptor for transforming growth factor-β (TGF-β). VEGF, PLGF, TGF-β as well as other pro-angiogenic proteins are known to be essential for normal placental and fetal vascular development. Soluble Flt-1 and sENG are almost undetectable in the circulation of non-pregnant individuals, and are produced in large quantities by the placenta leading to marked elevation in their circulating levels during pregnancy which steadily rise until term (195, 196). These two soluble receptors are increased in serum and placentas of preeclamptic women compared to normal pregnancies and their abnormal elevation presages the development of preeclampsia. Experimental evidence indicates that sENG cooperates with sFlt-1 to induce endothelial dysfunction in vitro and preeclampsia in vivo (197). It is thought that sFlt-1 and sENG neutralize their ligands, reducing the concentration of VEGF, PLGF, and TGF-b in maternal circulation, which results in a shift in the angiogenic balance towards anti-angiogenesis, which in turn leads to endothelial damage and the clinical onset of the syndrome. However, large prospective studies have failed to show sufficient accuracy of these biomarkers for clinical utility in preeclampsia prediction (198, 199).

PLACENTAL PEPTIDE HORMONES: OTHER PLACENTAL PEPTIDES

In addition to the pregnancy-related proteins produced analogous to hypothalamic and pituitary glycoproteins, the placenta also produces several other proteins that have no known analogues in the non-pregnant state. These proteins have been isolated and identified from serum drawn during pregnancy or purified from placental tissue. Figure 11 shows the changes in concentration of each of these pregnancy-related proteins throughout gestation.

Placental Pregnancy-Specific b1-Glycoprotein (SP1)

Pregnancy-specific b1-glycoprotein is a glycoprotein hormone that can be detected about 18-23 days after ovulation. It is secreted from trophoblast cells (200, 201). Initially, it exhibits a 2- to 3-day doubling time, reaching peak concentrations between 100-200 ng/mL at term. Pregnancy-specific b1-glycoprotein has immunosuppressive effects on T-lymphocyte proliferation, and is thought to be involved in preventing rejection of the implanting conceptus (202).

Placental Pregnancy-Associated Plasma Protein-A (PAPP-A)

Pregnancy-associated plasma protein-A is the largest of the pregnancy-related glycoproteins. It originates, mainly, from placental syncytiotrophoblasts (203, 204). Pregnancy-associated plasma protein-A can first be detected at approximately 32-33 days after ovulation. With a 3-day doubling time, its levels initially rise rapidly, and then continue to rise more slowly until term (203). Like SP-1 and hCG, PAPP-A is believed to play an immunosuppressive role in pregnancy (204). It has recently gained favor as a clinically useful, first-trimester screening marker for Down syndrome (trisomy 21). Authors have confirmed decreased PAPP-A levels in association with early pregnancy failure (205). However, when compared with serum hCG and progesterone measurements to evaluate the clinical usefulness of PAPP-A values in predicting the outcome of early pregnancy, hCG and progesterone remained the best clinical tools (206).

Placental Protein-5 (PP5)

This glycoprotein is produced by the syncytiotrophoblasts. It is detected beginning at 42 days after ovulation, and steadily rises until term (207). Placental protein-5 has anti-thrombin and anti-plasmin activities, and is believed to be a naturally occurring blood coagulation inhibitor active at the implantation site (208).

Figure 11. Maternal serum concentrations of human chorionic gonadotropin (hCG) and some other pregnancy-associated protein hormones (SP-1, PAPP-A, PP-5) throughout pregnancy. The timing of implantation, missed menses and parturition is shown to demonstrate the temporal relationships. (Modified from (209), with permission)

PLACENTAL METABOLIC PROTEINS  

Placental Leptin  

Leptin is a key regulator of satiety and body mass index (BMI), and its levels are thought to reflect the amount of energy stores and nutritional state (210).  The placenta is the principal source of leptin during pregnancy (211). Most of the leptin produced by the placenta is secreted into the maternal circulation, and as a consequence leptin levels are elevated during pregnancy.  In the first trimester, maternal plasma leptin levels are double nonpregnant values and continue to increase during the second and third trimesters (212-214).  In the second and third trimesters leptin is also expressed in the chorion and amnion (215).  The amount of leptin directed to the fetus is uncertain, and its role in fetal development is also unclear.  Leptin levels decline to normal nonpregnant levels within 24 hours after delivery (216). Interestingly, leptin levels during pregnancy do not correlate with BMI as they do in the nonpregnant state (217).  Although not clear, it is thought that leptin may be utilized by the placenta to modulate maternal metabolism and partition energy supplies to the fetus (218). There is evidence that placental leptin is anti-apoptotic and promotes proliferation, protein synthesis and the expression of tolerogenic maternal response molecules such as HLA-G (219). Placental leptin expression is regulated by hCG, insulin, steroids, hypoxia and many other growth hormones, suggesting that it may have an important endocrine function in trophoblast cells (219). Additionally, the human placenta also expresses leptin receptors, and therefore can act in a paracrine manner to modulate placental function (220, 221).

Placental Ghrelin   

Ghrelin, is a gastric peptide isolated primarily from the stomach which is thought to stimulate GH release and participates in the regulation of energy homeostasis, increasing food intake, decreasing energy output, as well as exert a lipogenetic effect (222).  Ghrelin and its receptors have been isolated in the placenta, clearly indicating a role for ghrelin in reproduction. Circulating ghrelin levels peak at mid-gestation, then with advancing gestational age declining ghrelin levels are observed.  After delivery, near prepregnancy levels of ghrelin are seen (223).  It is thought that ghrelin may well be involved in regulation of energy intake during pregnancy (224), however its exact role is still unknown.

PLACENTAL MATURATION

As pregnancy advances, the relative numbers of trophoblasts increase as feto-maternal exchange begins to dominate the placenta's secretory functions. Later, throughout the second and third trimester, the placenta adapts its structure to reflect its function such that near term, the villi consist mainly of fetal capillaries with sparse supporting stroma beyond that which is required to maintain its anatomic integrity. In contrast to the early placental villus where trophoblasts are abundant as part of a continuous layer of basal cytotrophoblasts, the term placenta's membranous interface between the fetal and maternal circulation is extremely thin (65). Thus, as the gestation progresses toward term, the number of cytotrophoblasts declines and the remaining syncytial layer becomes thin and barely visible. This structural arrangement facilitates transport of compounds across the feto-maternal interface. Consistent with the cytologic changes that occur in the maternal fetal interface from mid-gestation to term, striking changes in the global gene expression profile of this tissue has been demonstrated over this interval (225).

FETAL COMPARTMENT

The endocrine system, a system that is functional from the time of intrauterine existence through old age, is one of the first systems to develop during fetal life. As in the placenta, the regulation of the fetal endocrine system relies, to some extent, on precursors secreted by the other compartments. As the fetus develops, its endocrine system matures and eventually becomes more independent, preparing the fetus for extrauterine life.

Fetal Hypothalamus and Pituitary

By the end of the fifth week of gestation, the primitive hypothalamus can be identified as a swelling on the inner surface of the diencephalic neural canal (226).  By the 9th to 10th week, the median eminence of the hypothalamus is evident.  By week 14 to 16 the hypophysiotropic hormones GnRH, TRH, CRH, GHRH and somatostatin appear in the fetal hypothalamus (227) .  The portal-vessel system that delivers the releasing hormones to the anterior pituitary is fully developed by 18 weeks of gestation (227).

The anterior pituitary cells that develop from those cells lining Rathke's pouch are capable of secreting growth hormone (GH), follicle-stimulating hormone (FSH), luteinizing hormone (LH) and adrenocorticotropic hormone (ACTH), in vitro, as early as 7 weeks of fetal life (Figure 12).

Figure 12. Fetal serum pituitary hormone levels. PrL indicates prolactin; TSH, thyroid-stimulating hormone; ACTH, corticotropin; GH, growth hormone; LH/FSH, luteinizing hormone/follicle stimulating hormone. (Modified from (228), with permission)

Fetal Thyroid Gland

The fetal thyroid gland develops initially in the absence of detectable TSH. By 12 weeks’ gestation, the thyroid is capable of iodine-concentrating activity and thyroid hormone synthesis (226) .  Prior to that time, the maternal thyroid appears to be the primary source for T4.  The levels of TSH and T4 are relatively low in fetal blood until mid-gestation. At 24-28 weeks' gestation, serum T4 and reverse tri-iodothyronine (rT3) concentrations begin to rise progressively until term while the TSH concentration peaks. At birth, there is an abrupt release of TSH, T4, and T3. The relative hyperthyroid state of the newborn is believed to facilitate thermoregulatory adjustments for extrauterine life.  The function of the fetal thyroid hormones is crucial for somatic growth and neonatal adaptation. 

Fetal Gonads

The internal genitalia in the embryo have the inherent tendency to feminize. The Wolffian (mesonephric) and Mullerian (paramesonephric) ducts are discrete primordia that temporarily coexist in all embryos during the ambisexual undifferentiated development period (up to 8 weeks). The critical factors in determining which of the duct structures stabilize or regress are the hormones secreted by the testes: Anti-Mullerian hormone (AMH) and testosterone. The testis is histologically identifiable at 6 weeks’ gestation. Primary testis differentiation begins with development of the Sertoli cells at 8 weeks’ gestation. SRY, the sex-determining region on the Y chromosome, determines male gonadal sex and directs the differentiation of the Sertoli cell (229).  Sertoli cells secrete AMH which triggers the resorption of the Mullerian ducts in males and prevents development of female internal structures (230). At approximately 8 weeks’ gestation Leydig cells differentiate and testosterone secretion commences.  Maximum levels of fetal testosterone are observed at about 15 – 18 weeks and decrease thereafter.

Differentiation of the ovaries occurs several weeks later than that of the testis.  If the primordial germ cells lack the SRY region on the Y chromosome, ovaries develop from the indifferent gonads.  Fetal ovarian function becomes apparent by 7 to 8 weeks gestation; the time when the ovary becomes morphologically recognizable. During this time ovarian differentiation is occurring with mitotic multiplication of germ cells, reaching 6-7 million oogonia, their maximal number, by 16-20 weeks’ gestation (231, 232). 

The pattern of luteinizing hormone (LH) levels in fetal plasma parallels that of follicle-stimulating hormone (FSH). The decline in pituitary gonadotropin content, and plasma concentration of gonadotropins after mid-gestation is believed to result from the maturation of the hypothalamic-pituitary-gonadal axis. The hypothalamus becomes progressively more sensitive to sex steroids originating from the placenta and circulating in fetal blood. Early secretion of fetal testosterone is important in initiating sexual differentiation in males. In the absence of testosterone, the Wolffian system regresses. Human chorionic gonadotropin (hCG), supplemented by fetal LH, is believed to be the primary stimulus effecting the early development and growth of Leydig cells as well as stimulating the subsequent peak of testosterone production. In females, the fetal ovary is involved primarily in the formation of follicles and germ cells and less involved in hormone secretion.

Fetal Adrenal Glands

The human fetal adrenal gland is a remarkable organ due to its incredible capacity for steroid biosynthesis in utero, and because of its unique morphologic features. The human fetal adrenals are disproportionately large, and at mid-pregnancy their size exceeds that of the fetal kidneys. At term, the adrenals are as large as those of adults, weighing 10 grams or more. The region that ultimately develops into the adult adrenal cortex, the outer or definitive zone, accounts for only about 15% of the fetal gland (Figure 13). The unique inner or fetal zone comprises 80-85% of the volume of the adrenal in utero, and is largely responsible for the tremendous secretory capacity of this organ. The fetal zone rapidly undergoes involution at parturition and by one year it has completely disappeared (233). Changes in the fetal adrenal volume throughout fetal life and into young adulthood are graphically depicted in Figure 14.

The adrenal function of 10 preterm infants of gestational age 27-34 weeks was assessed for up to 80 days after delivery. The changes in steroid excretion with time in preterm infants of gestation over 28 weeks reflect involution of the fetal adrenal zone at a similar rate to term infants. These findings are consistent with the removal at birth of the inhibitory effects of estrogen on the 3 beta-hydroxysteroid dehydrogenase enzyme. The continued function of the adrenal fetal zone beyond the first month in preterm infants of less than 28 weeks’ gestation may however be due to persistence of some other fetal regulatory adrenal mechanism. This suggests that it is gestation that determines fetal zone activity rather than birth (234).

The fetal adrenal gland secretes large quantities of steroid hormones (up to 200-mg daily) near term. The rate of steroidogenesis is 5-times that observed in the adrenal glands of adults at rest. The principal steroids secreted are C-19 steroids (mainly DHEAS), which serve as substrates for estrogen biosynthesis by the placenta (Figure 13).

The fetal adrenal gland contains a zone, unique to in-utero fetal life that accounts for the rapid growth of the adrenal gland; this zone regresses during the first few weeks after birth. In addition to the fetal zone, an outer layer of cells forms the adrenal cortex (definitive zone). The fetal zone differs not only histologically, but also biochemically from the cortex (i.e., the fetal zone is deficient in 3b-hydroxysteroid dehydrogenase enzyme activity and, therefore, secretes C-19 steroids (mainly DHEAS); the cortex secretes primarily cortisol).

Figure 14. An illustration demonstrating generalized pathways for steroid hormone formation in the fetal adrenal gland. DHA: dehydroepiandrosterone. DHAS: dehydroepiandrosterone sulfate. LDL: low-density lipoprotein cholesterol. (Modified from (235), with permission)

Figure 15. Changes in the fetal adrenal volume throughout fetal life and into young adulthood. (Modified from (236), with permission)

Research involving the fetal adrenal gland has attempted to determine the factors that stimulate and regulate fetal adrenal growth and steroidogenesis. Other work has focused on the mechanisms responsible for fetal zone atrophy after delivery. All investigations have shown that, in vitro, adrenocorticotropic (ACTH) stimulates steroidogenesis. Furthermore, there is clinical evidence that, in vivo, ACTH is the major trophic hormone of the fetal adrenal gland. For example, in anencephalic fetuses, the plasma levels of ACTH are very low and the fetal zone is markedly atrophic. Maternal glucocorticoid therapy suppresses fetal adrenal steroidogenesis by suppressing fetal ACTH secretion. Despite these observations, ACTH -related peptides, growth factors and other hormones have been proposed as possible trophic hormones for the fetal zone. After birth, the adrenal gland shrinks in size by more than 50% because of the regression of fetal zone cells.

Fetal Parathyroid Glands and Calcium Homeostasis

In the fetus, calcium concentrations are regulated by the movement of calcium across the placenta from the maternal compartment. In order to maintain fetal bone growth, the maternal compartment undergoes adjustments that provide a net transfer of sufficient calcium to the fetus. Maternal compartment changes that permit calcium accumulation include increases in maternal dietary intake, increases in maternal 1, 25-dihydroxyvitamin D3 levels, and increases in parathyroid hormone (PTH) levels.  The levels of total calcium and phosphorus decline in maternal serum, but ionized calcium levels remain unchanged. During pregnancy, the placenta forms a calcium pump in which a gradient of calcium and phosphorus is established which favors the fetus.  Thus, circulating fetal calcium and phosphorus levels increase steadily throughout gestation. Furthermore, fetal levels of total and ionized calcium, as well as phosphorus, exceed maternal levels at term.

By 10-12 weeks' gestation, the fetal parathyroid glands secrete PTH. Fetal plasma levels of PTH are low during gestation, but increase after delivery. In contrast to the unchanged maternal calcitonin levels, the fetal thyroid gland produces increasing levels of calcitonin. Since there is no transfer of parathyroid hormone across the placenta, changes noted in fetal calcium levels are related to fetal changes in these hormones (PTH and calcitonin). These adaptations are consistent with the need to conserve calcium and stimulate bone growth within the fetus. After birth, neonatal serum calcium and phosphorus levels fall. Parathyroid hormone levels start to rise within 48 hours after birth. Calcium and phosphorus levels steadily increase over the following several days, with some dependence on dietary intake of milk.

Fetal Endocrine Pancreas

The pancreas’ exocrine function begins after birth, while the endocrine function (hormone release) can be measured from 10 to 15 weeks onward.  The α-cells which contain glucagon, and the β-cells which contain somatostatin, can be recognized by 8 weeks’ gestation (234). Alpha cells are more numerous than β-cells in the early fetal pancreas and reach a peak at midgestaion; β-cells increase through the second half of gestation so that by term the ratio of α-cells to beta cells is approximately 1:1 (237, 238).  Human pancreatic insulin and glucagon concentrations increase with advancing fetal age, and are higher than concentrations found in the adult pancreas. In vivo studies of umbilical cord blood obtained at delivery and fetal scalp blood samples obtained at term show that fetal insulin secretion is low and tends to be relatively unresponsive to acute changes in glucose. In contrast, fetal insulin secretion in vitro is responsive to amino acids and glucagon as early as 14 weeks' gestation. In maternal diabetes mellitus, fetal islet cells undergo hypertrophy such that the rate of insulin secretion increases.

Fetal Alpha-Fetoprotein (AFP)

Alpha-fetoprotein is a glycoprotein synthesized first by the yolk sac, then the gastrointestinal tract, and lastly by the fetal liver (239, 240). After entering the fetal urine, it is readily detected in amniotic fluid. Amniotic fluid AFP (afAFP) peaks between 10-13 weeks’ gestation, and then declines from 14-32 weeks. In the fetus, AFP peaks at 12-14 weeks and steadily decreases until term (241). The fall in fetal plasma AFP (fpAFP) is most likely due to the combination of increasing fetal blood volume and a decline in fetal production. The concentration gradient between fpAFP and maternal serum AFP (msAFP) is approximately 150- to 200-fold. Detectable as early as 7 weeks' gestation, msAFP reaches peak concentrations between 28-32 weeks (241). The seemingly paradoxical rise in msAFP in association with decreasing afAFP and fetal serum levels can be accounted for by the increasing placental permeability to fetal plasma proteins that occurs with advancing gestational age (241). Alpha-fetoprotein acts as an osmoregulator to help adjust fetal intravascular volume (241). It may also be involved in certain immunoregulatory functions (242). Amniotic fluid AFP and maternal serum AFP are clinically important because they are elevated in association conditions such as neural tube defects (243). Additionally, msAFP is decreased in pregnancies in which the fetus has Down syndrome (trisomy 21) (244).

MATERNAL COMPARTMENT

Maternal Hypothalamus and Pituitary

Little information is definitively known about the endocrine alterations of the maternal hypothalamus during pregnancy. Thought to result from estrogen stimulation, the anterior pituitary undergoes a 2- to 3-fold enlargement during pregnancy, primarily because of hyperplasia and hypertrophy of lactotroph cells. Thus, plasma prolactin levels parallel the increase in pituitary size throughout gestation. In contrast to the lactotrophs, the size of the other pituitary cells decreases or remains unaltered during pregnancy. In line with these findings, maternal levels of somatotrophs and gonadotrophs are lower and the level of thyrotrophs and corticotrophs remains unchanged.  In contrast, adrenocorticotrophic hormone (ACTH) levels do increase with advancing gestation. Corticotrophin-releasing hormone (CRH) in the maternal plasma increases during pregnancy due to increased placental secretion, but alterations in binding-protein concentrations prevent increased biologic activity of this releasing hormone.

The size of the posterior pituitary gland diminishes during pregnancy (245).  Additionally, maternal plasma arginine vasopressin (AVP) levels remain low throughout gestation and are not believed to play a pivotal role in human pregnancy.  In contrast, maternal oxytocin levels progressively increase in the maternal blood and parallel the increase in maternal serum levels of estradiol and progesterone (246). Uterine oxytocin receptors also increase throughout pregnancy, resulting in a 100 fold increase in oxytocin binding at term in the myometrium (247).

Maternal Thyroid Gland

As a result of increased vascularity and glandular hyperplasia, the thyroid gland increases in size by 18% during pregnancy; however, true goiter is not usually present (248).  Enlargement is associated with an increase in the size of the follicles with increased amounts of colloid and enhanced blood volume.  This enlargement may be a response to the thyrotropic effect of hCG, which may account for some of the increase in serum thyroglobulin concentrations observed during pregnancy.  During gestation the mother remains in a euthyroid state. Total thyroxine (T4) and tri-iodothyronine (T3) levels increase but do not result in hyperthyroidism because there is a parallel increase in T4-binding globulin that results from estrogen exposure (Figure 15). The increase seen in binding-protein concentrations is similar to that observed in women who use oral contraceptives (OC). A modest increase in the basal metabolic rate (BMR) rate occurs during pregnancy secondary to increasing fetal requirements. Some T4 and T3, but no TSH, are transferred across the placenta.

Figure 15. Relative changes in maternal thyroid function during the course of human pregnancy from fertilization to term. (Modified from (249), with permission)

Maternal Adrenal Glands

The maternal adrenal gland does not change morphologically during pregnancy.  However, plasma adrenal steroid levels increase with advancing gestation. Total plasma cortisol concentrations increase to three times nonpregnant levels by the third trimester.  The hypoestrogenic state of pregnancy results in increased hepatic production of cortisol-binding globulin. This increase in cortisol-binding globulin results in decreased metabolic clearance of cortisol, resulting in an increase in plasma free cortisol and total free cortisol.  Additionally, cortisol production increases due to an increase in maternal plasma ACTH concentration and the hyperresponsiveness of the adrenal cortex to the ACTH stimulation (250).  Despite the elevated free cortisol levels, pregnant women do not exhibit any overt signs of hypercortisolism, likely due to the anti-glucocorticoid activities of the elevated levels of progesterone.

Plasma renin substrate levels are increased as a consequence of the effects of estrogen on the liver.  The higher levels of renin and angiotensin during pregnancy, lead to elevated angiotensin II levels and markedly elevated levels of aldosterone.  Similar to cortisol, the elevated aldosterone levels do not have a detrimental effect on maternal health.  The high level of progesterone is thought to displace aldosterone from its renal receptors.

Androgen levels are elevated during pregnancy, but the free androgen levels remain normal to low secondary to the estrogen-induced increase in hepatic synthesis of sex hormone-binding globulin. Dehydroepiandrosterone (DHEA) and DHEAS production is increased twofold during pregnancy. However, serum concentrations of DHEAS are reduced to less than nonpregnant levels secondary to enhanced 16 –hydroxylation and placental use of DHEAS in estrogen production (251).

Maternal Endocrine Pancreas

A dual-hormone secretion mechanism is partially responsible for the metabolic adaptation of pregnancy in which glucose is spared for the fetus by the maternal endocrine pancreas. Compared to the non-pregnant state, in response to a glucose load, there is a greater release of insulin from the β-cells and a greater suppression of glucagon release from the α-cells. Associated with the increased release of insulin, the maternal pancreas undergoes β-cell hyperplasia and islet-cell hypertrophy, with an accompanying increase in blood flow to the endocrine pancreas. During pregnancy, when fasting blood glucose levels fall, they rise to a greater extent in response to a glucose load than do levels in non-pregnant women. The increased release of insulin is related to insulin resistance due to hPL, which spares transfer of glucose to the fetus. Glucagon levels are also suppressed in response to a glucose load, with the greatest suppression occurring near term.

REGULATION OF FETO-MATERNAL STEROIDOGENESIS

Using in vitro investigations utilizing placental tissue explants as well as in vivo, catheterized primate models to study steroidogenic regulation in pregnancy, researchers have determined LDL-cholesterol, fetal pituitary hormones, intra-placental regulators, and intra-adrenal regulators act as the primary modulators of feto-placental steroid production (252-254).

Regulation by Low Density Lipoprotein Cholesterol (LDL)

A limiting factor in adrenal steroid output is the availability of LDL-cholesterol, the primary lipoprotein used in fetal adrenal steroid synthesis (Figure 16). Circulating LDL-cholesterol accounts for 50-70% of the cholesterol utilized for fetal adrenal steroidogenesis (255-257). The fetal adrenal is known to contain high affinity, low-capacity LDL binding sites. The presence of ACTH increases this binding capacity (256, 258, 259). Within the adrenal gland, hydrolysis of LDL makes cholesterol available for conversion to steroids. The majority of fetal LDL-cholesterol is made, de novo, in the fetal liver (260). In addition, cortisol from the fetal adrenal cortex and estradiol (aromatized from fetal DHEAS) augment this de novo synthesis within the fetal liver. These systems interact in a manner that is linked, self-perpetuating, and serves to increase steroid production to meet the needs of the maturing fetus (260).

Figure 16. Shown are the maternal, placental and fetal compartments for estrogen and progesterone synthesis in human pregnancy. The fetal adrenal gland lacks 3β-hydroxysteroid dehydrogenase, but has sulfation and 16α-hydroxylase capabilities. Likewise, the placenta lacks 17α-hydroxylase activity but contains sulfatase in order to cleave the sulfated fetal products. Modified from (261), with permission)

Regulation by Fetal Pituitary Hormones  

Fetal ACTH regulates steroidogenesis in both adrenal zones. Adrenocorticotropic hormone receptor activity is diminished in the fetal zone of the cortex during the early second trimester when other factors, such as hCG, are more important in the maintenance of this zone (260). In vitro studies in human fetal adrenal tissue, demonstrate that ACTH stimulates the release of D5 pregnenolone sulfate and DHEAS, whereas in adult adrenal cortex secretes only cortisol when stimulated by ACTH (260). Moreover, ACTH can act on its own adrenal-cell membrane receptor to express a direct stimulatory effect on steroidogenic enzymes (260).

Adrenocorticotropic hormone extracted from the human fetal pituitary gland has been shown, in vitro, to stimulate the production of DHEAS and cortisol (262, 263). Interestingly, concentrations of ACTH throughout gestation do not correlate with the increasing mass of the fetal adrenal cortex or the increasing steroidogenic function that are hallmarks of the third trimester (259). Fetal pituitary ACTH is detectable by 9 weeks’ gestation (263, 264). Thereafter, levels of ACTH increase steadily until 20 weeks’ gestation. The levels remain stable until approximately 34 weeks, when a significant decline is initiated and persists until term (259).

Prolactin may act as a co-regulator, along with ACTH, hCG and certain growth factors, in fetal adrenal steroid production (265, 266). Both in vitro and in vivo, prolactin augments ACTH-stimulated adrenal androgen production (253). Fetal pituitary prolactin is detectable at 10 weeks’ gestation (264). Umbilical cord prolactin levels increase with advancing gestational age and rise in parallel with increased fetal adrenal mass (267).

Regulation by Intra-Placental Mechanisms  

The placenta is an important co-regulator of the fetal adrenal zone due its ability to secrete hCG, placental CRH, progesterone and estradiol (233). In vitro and in vivo, hCG receptor activity is present in the fetal zone, and hCG stimulates fetal adrenal production of DHEAS (233, 268). However, after the 20th week of gestation ACTH primarily influences the fetal zone of the adrenal, and at this time hCG plays only a minor role. Placental CRH, acts in a paracrine relationship with placental ACTH, to complement the actions of the fetal hypothalamus and pituitary in producing the surge in fetal glucocorticoids notable in the late third trimester as fetal growth and maturity become increasingly important (125, 269).

Placental progesterone inhibits D5 to D4 steroid transformations in the fetal zone of the adrenal (101, 270). This effect is another explanation for fetal adrenal 3β-HSD deficiency. Placental estradiol modifies the production and metabolism of corticosteroids and progesterone. In vivo, the placenta regulates the inter-conversion of maternal cortisol to cortisone, and the fetal pituitary production of ACTH (264, 269). Modulation of the transfer of maternal cortisol across the placenta, into the fetus, is the primary mechanism through which this effect occurs.

Regulation by Intra-Adrenal Mechanisms  

With advancing gestational age, the fetal adrenal becomes more sensitive to circulating ACTH (253). Between 32 and 36 weeks’ gestation, the fetal adrenal mass increases (271-273). Blood flow to the fetal adrenal is affected by many factors that, in turn, affect the exposure of the fetal adrenal receptors of the different trophic stimuli. Growth factors modulate adrenal steroid pathways just as they do in the adult adrenal cortex. The fetal adrenal produces IGF-I and IGF-II; ACTH originating from either the fetal pituitary or the placenta can stimulate production of their respective mRNAs (274, 275).

PARTURITION

Parturition is a coordinated process of transition from a quiescent myometrium to an active rhythmically contractile state requiring elegant interplay between placental, fetal and maternal compartments. Though fetal maturity does not always predate the onset of labor, the two processes are related in primates. The timing of birth is a crucial determinant of perinatal outcome. Both preterm birth (<37wk) and post-term pregnancy (>42 wk) are associated with greater risk of adverse perinatal outcomes. The traditional dogma, supported by robust evidence from animal studies, has the fetoplacental unit as being in charge of the timing of parturition (276). While this is certainly true in some species, the presence of such a “placental clock” is not established in humans. Rather, it has become clear that the maternal endometrium/decidua also plays an important role in triggering the cascade of event leading to parturition (277).  

The precise mechanisms involved in human parturition are thought to involve CRH, functional progesterone withdrawal, increased estrogen bioavailability, and finally, increased responsiveness of the myometrium to prostaglandins and oxytocin. There is no simple chain of events as there are in other species.

Numerous lines of evidence support a role for CRH in human parturition. Studies have demonstrated increased CRH and decreased CRH-binding protein levels prior to the onset of both term and preterm labor (278, 279). CRH directly stimulates release of prostaglandins in decidua and myometrium (280). Interestingly, a paradoxical augmentation of placental CRH release by serum cortisol is maximal in the last ten weeks of pregnancy. This may be a function of cortisol competition with progesterone for placental glucocorticoid receptors, thereby blocking the inhibitory action of progesterone on CRH synthesis (281).

The ratios of estradiol and progesterone in various animal models are closely related to the stimulation of myometrial gap-junction formation (282). With decreasing progesterone relative to estradiol, gap junctions permit cell-cell communication for the synchronized myometrial smooth muscle contractions required for labor. Progesterone and the estrogens are antagonistic in the parturition process. Progesterone produces uterine relaxation, stabilizing lysosomal membranes and inhibiting prostaglandin synthesis and release. By contrast, estrogens destabilize lysosomal membranes and augment the synthesis of prostaglandin and their release (283). Although gradual increase in umbilical cord DHEAS and maternal estriol occurs toward term, there is no corresponding drop in either fetal or maternal progesterone concentrations (284).

Though a reduction in maternal or fetal progesterone levels during spontaneous labor has not been documented, functional progesterone withdrawal at the receptor level is believed to be involved in the process of parturition. This may occur via altered progesterone receptor isoform PR-A/PR-B levels in myometrium (285). Undoubtedly, progesterone is important in uterine quiescence because in the first trimester removal of the corpus luteum leads rapidly to myometrial contractions (84). Likewise, labor ensues following the administration of progesterone receptor antagonists in the third trimester (286). The anti-progesterone agents occupy progesterone receptors and inhibit the action of progesterone, which is clearly essential for maintenance of uterine quiescence. Consistent with these findings, pharmacologic treatment of women at risk for preterm labor with progesterone or synthetic progestational agents has demonstrated efficacy in the prevention of preterm labor (287-289).

A role for estrogen in the process of parturition is supported by the finding that pregnancies are often prolonged when estrogen levels in maternal blood and urine are low, as in placental sulfatase deficiency or when associated with anencephaly (290). In human studies, there is a correlation in uterine activity with circulating maternal estrogens and progesterone as labor approaches (291-293). Feto-placental estrogens are closely linked to myometrial irritability, contractility, and labor. In primates, estrogens ripen the cervix, initiate uterine activity, and established labor (294). Estrogens also increase the sensitivity of the myometrium to oxytocin by augmenting prostaglandin biosynthesis (283, 295). Because placental release of estrogens is linked to the fetal hypothalamus, pituitary, adrenals, and placenta the fetal pituitary adrenal axis appears to fine-tune parturition timing in part through its effect on estrogen production.

Prostaglandins (PG) are thought to play a central role in human parturition. For years, it has been known that rupture, stripping, or infection of the fetal membranes, as well as instillation of hypertonic solutions into the amniotic fluid results in the onset of labor. These facts have led to the hypothesis that a fetal-amniotic fluid-fetal membrane complex is a metabolically active unit that triggers the onset of labor. Evidence supporting a causative role of prostaglandins in the labor process is present since PGs induce myometrial contractions in all stages of gestation. While there is still no direct evidence relating endogenous PGs to labor,  there are several lines of evidence implicating PGs in this process; PG levels increase in maternal circulation and amniotic fluid in association with labor; indomethacin prevents the onset of labor in nonhuman primates and stops preterm labor in humans; stimuli that are known to induce labor (e.g. cervical ripening, rupture of membranes) are associated with PG release; the process of cervical ripening is mediated by PGs. Important to this hypothesis is the understanding that at least one mechanism in the onset of parturition is the release of stored precursors of PGs from the fetal membranes.

The major precursor for PGs is arachidonic acid, which is stored in glycerophospholipids. The fetal membranes are enriched with two major glycerophospholipids, phosphatidylinositol and phosphatidylethanolamine. As gestation advances, the progressively increasing levels of estrogen stimulate the storage, in fetal membranes, of these glycerophospholipids containing arachidonic acid.

A series of fetal membrane lipases, including phospholipase A2 and Phospholipase C control the release of arachidonic acid from storage in fetal membrane phospholipids. Once in a free state, arachidonic acid is available for conversion to PG. Additional factors that augment and accentuate the normal process of labor include the liberation of corticosteroid by the mother and fetus, resulting in a decrease in the production of myometrial prostacyclin, a smooth muscle relaxant.

Active labor is characterized by a dramatic increase in the number of oxytocin receptors in the myometrium. Once begun, the process appears to be self-perpetuating. The level of maternal catecholamines increases, resulting in the liberation of free fatty acids, including arachidonic acid; there is also an increase in the level of maternal or fetal cortisol, which decreases the production of uterine smooth muscle prostacyclin. It is unlikely that oxytocin is the initiator of labor despite the fact that oxytocin receptors are present in the myometrium and increase before labor, and it stimulates decidual prostaglandin E2 and prostaglandin F2a production. There is firm evidence of increasing, rhythmical fetal adrenal and placental steroid output over the 5 weeks just before term that is important in preparing human pregnancy for the final cascade of oxytocin and prostaglandins that stimulate labor (283, 291-293, 295, 296).

KEY POINTS

  • Synchrony between the development of the early embryo and establishment of a receptive endometrium is necessary to allow implantation and subsequent progression of pregnancy.
  • The placenta is a unique, dynamic organ with the inherent ability to produce, regulate, and inhibit factors that directly affect fetal growth and development.
  • During the luteal-placental transition period, between 6-10 weeks of gestation, corpus luteal function and progesterone production naturally declines and shifts to the developing placenta.
  • Steroidogenesis in pregnancy is characterized by enzymatic deficiencies within the placental and fetal compartments which foster interdependent transfer of precursors among compartments for the synthesis of steroid hormones. This process is modulated by LDL-cholesterol, fetal pituitary hormones, intra-placental regulators, and intra-adrenal regulators.
  • Redundancy in protein hormone – receptor interactions such as hPL and hPGH serve to ensure that adequate nutrition is supplied to the developing fetus.
  • A relatively insulin resistant state is generated within the maternal compartment to supply glucose and free fatty acids for fetal nutrition.
  • Human parturition exemplifies the interplay between placental, fetal, and maternal compartments, characterized by increased estrogen bioavailability, functional progesterone withdrawal, increased CRH synthesis and release, culminating in increased responsiveness of the myometrium to prostaglandins and oxytocin.

ACKNOWLEDGMENT

In addition to the journal and text references listed above, the following sources were used in the preparation of this chapter:

Taylor HS, Pal L, Seli E (eds.). Speroff’s Clinical Gynecologic Endocrinolofy & Infertility. Ninth edition, 2020. Wolters-Kluwer, Philadelphia.

Gabbe SG, Niebyl JR, Simpson JL [eds.]. Obstetrics: normal and problems pregnancies. Fifth edition, 2007. Churchill-Livingstone, New York.
Benirschke K, Kaufmann P, Baergen RN [eds.]. Pathology of the human placenta. Fifth edition, 2006. Springer, New York.Strauss JF,
Barbieri RL [eds.]. Yen and Jaffe’s Reproductive endocrinology: physiology, pathophysiology and clinical management. Fifth edition, 2004. Elsevier Saunders, Philadelphia.
Reece EA, Hobbins JC [eds.]. Clinical obstetrics: the fetus and mother. Third edition, 2007. Wiley-Blackwell, Malden, MA.

REFERENCES

  1. Mesino S: The Endocrinology of Human Pregnancy and Fetoplacental Neuroendocrine Development. ; in Yen & Jaffe Reproductive Endocrinology. Edited by Jaffe Y, 2009
  2. Strowitzki T, Germeyer A, Popovici R, et al.: The human endometrium as a fertility-determining factor. Human reproduction update 12:617-30, 2006
  3. Finn CA, Martin L: The control of implantation. Journal of reproduction and fertility 39:195-206, 1974
  4. Martin J, Dominguez F, Avila S, et al.: Human endometrial receptivity: gene regulation. Journal of reproductive immunology 55:131-9, 2002
  5. Gipson IK, Blalock T, Tisdale A, et al.: MUC16 is lost from the uterodome (pinopode) surface of the receptive human endometrium: in vitro evidence that MUC16 is a barrier to trophoblast adherence. Biology of reproduction 78:134-42, 2008
  6. Sharkey AM, Smith SK: The endometrium as a cause of implantation failure. Best practice & research Clinical obstetrics & gynaecology 17:289-307, 2003
  7. Chan RW, Schwab KE, Gargett CE: Clonogenicity of human endometrial epithelial and stromal cells. Biol Reprod 70:1738-50, 2004
  8. Taylor HS: Endometrial cells derived from donor stem cells in bone marrow transplant recipients. JAMA 292:81-5, 2004
  9. Tal R, Shaikh S, Pallavi P, et al.: Adult bone marrow progenitors become decidual cells and contribute to embryo implantation and pregnancy. PLoS Biol 17:e3000421, 2019
  10. Lima PD, Zhang J, Dunk C, et al.: Leukocyte driven-decidual angiogenesis in early pregnancy. Cellular & molecular immunology 11:522-37, 2014
  11. Hofmann AP, Gerber SA, Croy BA: Uterine natural killer cells pace early development of mouse decidua basalis. Molecular human reproduction 20:66-76, 2014
  12. Erlebacher A: Immunology of the maternal-fetal interface. Annu Rev Immunol 31:387-411, 2013
  13. DeMayo FJ, Lydon JP: 90 YEARS OF PROGESTERONE: New insights into progesterone receptor signaling in the endometrium required for embryo implantation. J Mol Endocrinol 65:T1-T14, 2020
  14. Lydon JP, DeMayo FJ, Funk CR, et al.: Mice lacking progesterone receptor exhibit pleiotropic reproductive abnormalities. Genes Dev 9:2266-78, 1995
  15. Mulac-Jericevic B, Lydon JP, DeMayo FJ, et al.: Defective mammary gland morphogenesis in mice lacking the progesterone receptor B isoform. Proc Natl Acad Sci U S A 100:9744-9, 2003
  16. Mulac-Jericevic B, Mullinax RA, DeMayo FJ, et al.: Subgroup of reproductive functions of progesterone mediated by progesterone receptor-B isoform. Science 289:1751-4, 2000
  17. Chwalisz K: The use of progesterone antagonists for cervical ripening and as an adjunct to labour and delivery. Human reproduction 9 Suppl 1:131-61, 1994
  18. Shaw KA, Topp NJ, Shaw JG, et al.: Mifepristone-misoprostol dosing interval and effect on induction abortion times: a systematic review. Obstetrics and gynecology 121:1335-47, 2013
  19. Lakha F, Ho PC, Van der Spuy ZM, et al.: A novel estrogen-free oral contraceptive pill for women: multicentre, double-blind, randomized controlled trial of mifepristone and progestogen-only pill (levonorgestrel). Human reproduction 22:2428-36, 2007
  20. Spitz IM, Croxatto HB, Lahteenmaki P, et al.: Effect of mifepristone on inhibition of ovulation and induction of luteolysis. Human reproduction 9 Suppl 1:69-76, 1994
  21. Giudice LC: Microarray expression profiling reveals candidate genes for human uterine receptivity. American journal of pharmacogenomics : genomics-related research in drug development and clinical practice 4:299-312, 2004
  22. Hsieh-Li HM, Witte DP, Weinstein M, et al.: Hoxa 11 structure, extensive antisense transcription, and function in male and female fertility. Development 121:1373-85, 1995
  23. Satokata I, Benson G, Maas R: Sexually dimorphic sterility phenotypes in Hoxa10-deficient mice. Nature 374:460-3, 1995
  24. Taylor HS, Arici A, Olive D, et al.: HOXA10 is expressed in response to sex steroids at the time of implantation in the human endometrium. The Journal of clinical investigation 101:1379-84, 1998
  25. Taylor HS, Igarashi P, Olive DL, et al.: Sex steroids mediate HOXA11 expression in the human peri-implantation endometrium. The Journal of clinical endocrinology and metabolism 84:1129-35, 1999
  26. Cakmak H, Taylor HS: Implantation failure: molecular mechanisms and clinical treatment. Human reproduction update 17:242-53, 2011
  27. Ochoa-Bernal MA, Fazleabas AT: Physiologic Events of Embryo Implantation and Decidualization in Human and Non-Human Primates. Int J Mol Sci 21, 2020
  28. Wilcox AJ, Baird DD, Weinberg CR: Time of implantation of the conceptus and loss of pregnancy. The New England journal of medicine 340:1796-9, 1999
  29. Morton H, Cavanagh AC, Athanasas-Platsis S, et al.: Early pregnancy factor has immunosuppressive and growth factor properties. Reproduction, fertility, and development 4:411-22, 1992
  30. Morton H, Rolfe BE, Cavanagh AC: Pregnancy proteins: basic concepts and clinical applications. Semin Reprod Endocrinol 10:72, 1992
  31. Cavanagh AC, Morton H, Rolfe BE, et al.: Ovum factor: a first signal of pregnancy? Am J Reprod Immunol 2:97-101, 1982
  32. Morton H, Rolfe BE, Cavanagh AC: Ovum factor and early pregnancy factor. Current topics in developmental biology 23:73-92, 1987
  33. Croxatto HB, Ortiz ME, Diaz S, et al.: Studies on the duration of egg transport by the human oviduct. II. Ovum location at various intervals following luteinizing hormone peak. American journal of obstetrics and gynecology 132:629-34, 1978
  34. Buster JE, Bustillo M, Rodi IA, et al.: Biologic and morphologic development of donated human ova recovered by nonsurgical uterine lavage. American journal of obstetrics and gynecology 153:211-7, 1985
  35. Macklon NS, Brosens JJ: The human endometrium as a sensor of embryo quality. Biology of reproduction 91:98, 2014
  36. Craciunas L, Gallos I, Chu J, et al.: Conventional and modern markers of endometrial receptivity: a systematic review and meta-analysis. Human reproduction update 25:202-23, 2019
  37. The morphological and functional development of the fetus. East Norwalk: Appleton & Lange, 1989
  38. Shutt DA, Lopata A: The secretion of hormones during the culture of human preimplantation embryos with corona cells. Fertility and sterility 35:413-6, 1981
  39. Laufer N, DeCherney AH, Haseltine FP, et al.: Steroid secretion by the human egg-corona-cumulus complex in culture. The Journal of clinical endocrinology and metabolism 58:1153-7, 1984
  40. Punnonen R, Lukola A: Binding of estrogen and progestin in the human fallopian tube. Fertility and sterility 36:610-4, 1981
  41. Hsueh AJ, Peck EJ, Jr., clark JH: Progesterone antagonism of the oestrogen receptor and oestrogen-induced uterine growth. Nature 254:337-9, 1975
  42. Ciarmela P, Islam MS, Reis FM, et al.: Growth factors and myometrium: biological effects in uterine fibroid and possible clinical implications. Human reproduction update 17:772-90, 2011
  43. Critchley HO, Brenner RM, Henderson TA, et al.: Estrogen receptor beta, but not estrogen receptor alpha, is present in the vascular endothelium of the human and nonhuman primate endometrium. The Journal of clinical endocrinology and metabolism 86:1370-8, 2001
  44. Albrecht ED, Robb VA, Pepe GJ: Regulation of placental vascular endothelial growth/permeability factor expression and angiogenesis by estrogen during early baboon pregnancy. The Journal of clinical endocrinology and metabolism 89:5803-9, 2004
  45. Albrecht ED, Aberdeen GW, Niklaus AL, et al.: Acute temporal regulation of vascular endothelial growth/permeability factor expression and endothelial morphology in the baboon endometrium by ovarian steroids. The Journal of clinical endocrinology and metabolism 88:2844-52, 2003
  46. Ma W, Tan J, Matsumoto H, et al.: Adult tissue angiogenesis: evidence for negative regulation by estrogen in the uterus. Molecular endocrinology 15:1983-92, 2001
  47. Bonduelle ML, Dodd R, Liebaers I, et al.: Chorionic gonadotrophin-beta mRNA, a trophoblast marker, is expressed in human 8-cell embryos derived from tripronucleate zygotes. Human reproduction (Oxford, England) 3:909-14, 1988
  48. Lopata A, Hay DL: The surplus human embryo: its potential for growth, blastulation, hatching, and human chorionic gonadotropin production in culture. Fertility and sterility 51:984-91, 1989
  49. Hay DL, Lopata A: Chorionic gonadotropin secretion by human embryos in vitro. The Journal of clinical endocrinology and metabolism 67:1322-4, 1988
  50. Enders AC: Embryo implantation, with emphasis on the rhesus monkey and the human. Reproduccion 5:163-7, 1981
  51. Tulchinsky D, Hobel CJ: Plasma human chorionic gonadotropin, estrone, estradiol, estriol, progesterone, and 17 alpha-hydroxyprogesterone in human pregnancy. 3. Early normal pregnancy. American journal of obstetrics and gynecology 117:884-93, 1973
  52. Chard T: Proteins of the human placenta: some general concepts; in Pregnancy Proteins: Biology, Chemistry and Clinical Application. Edited by Grudzinskas J, Teisner B, Sepala M. San Diego: Academic Press, 1982
  53. Saijonmaa O, Laatikainen T, Wahlstrom T: Corticotrophin-releasing factor in human placenta: localization, concentration and release in vitro. Placenta 9:373-85, 1988
  54. Khodr GS, Siler-Khodr TM: Placental luteinizing hormone-releasing factor and its synthesis. Science (New York, NY 207:315-7, 1980
  55. Shambaugh G, 3rd, Kubek M, Wilber JF: Thyrotropin-releasing hormone activity in the human placenta. The Journal of clinical endocrinology and metabolism 48:483-6, 1979
  56. Al-Timimi A, Fox H: Immunohistochemical localization of follicle-stimulating hormone, luteinizing hormone, growth hormone, adrenocorticotrophic hormone and prolactin in the human placenta. Placenta 7:163-72, 1986
  57. Hay DL: Placental histology and the production of human choriogonadotrophin and its subunits in pregnancy. British journal of obstetrics and gynaecology 95:1268-75, 1988
  58. Harada A, Hershman JM: Extraction of human chorionic thyrotropin (hCT) from term placentas: failure to recover thyrotropic activity. The Journal of clinical endocrinology and metabolism 47:681-5, 1978
  59. Steiner D: Peptide hormone precursors: biosynthesis, processing, and significance; in Peptide Hormones. Edited by Parson J. Baltimore: University Park Press, 1976
  60. Hoshina M, Hussa R, Pattillo R, et al.: The role of trophoblast differentiation in the control of the hCG and hPL genes. Advances in experimental medicine and biology 176:299-312, 1984
  61. Hoshina M, Boime I, Mochizuki M: [Cytological localization of hPL, hCG, and mRNA in chorionic tissue using in situ hybridization]. Nippon Sanka Fujinka Gakkai zasshi 36:397-404, 1984
  62. Kurman RJ, Young RH, Norris HJ, et al.: Immunocytochemical localization of placental lactogen and chorionic gonadotropin in the normal placenta and trophoblastic tumors, with emphasis on intermediate trophoblast and the placental site trophoblastic tumor. Int J Gynecol Pathol 3:101-21, 1984
  63. Kasai K, Aochi H, Shik SS, et al.: [Production and localization of human prolactin in the tissues associated with pregnancy (Report I) (author's transl)]. Nippon Naibunpi Gakkai zasshi 56:1574-80, 1980
  64. Watkins WB, Yen SS: Somatostatin in cytotrophoblast of the immature human placenta: localization by immunoperoxidase cytochemistry. The Journal of clinical endocrinology and metabolism 50:969-71, 1980
  65. Chard T, Grudzinskas JG: Pregnancy protein secretion. Semin Reprod Endocrinol 10:61, 1992
  66. Jones EE: Abnormal ovulation and implantation; in Medicine of the mother and fetus. Edited by Reece EA, Hobbins JC. Philadelphia: JB Lippincott Company, 1992
  67. Murphy BEP: Cortisol economy in the human fetus; in The Endocrine Function of the Human Adrenal Cortex. Edited by James VHT, Serio M, Gusti G. San Diego: Academic Press, 1978
  68. Murphy BE: Cortisol and cortisone in human fetal development. Journal of steroid biochemistry 11:509-13, 1979
  69. Handwerger S, Brar A: Placental lactogen, placental growth hormone, and decidual prolactin. Semin Reprod Endocrinol 10:106, 1992
  70. Maslar IA, Ansbacher R: Effects of progesterone on decidual prolactin production by organ cultures of human endometrium. Endocrinology 118:2102-8, 1986
  71. Raabe MA, McCoshen JA: Epithelial regulation of prolactin effect on amnionic permeability. American journal of obstetrics and gynecology 154:130-4, 1986
  72. Clements JA, Reyes FI, Winter JS, et al.: Studies on human sexual development. IV. Fetal pituitary and serum, and amniotic fluid concentrations of prolactin. The Journal of clinical endocrinology and metabolism 44:408-13, 1977
  73. Luciano AA, Varner MW: Decidual, amniotic fluid, maternal and fetal prolactin in normal and abnormal pregnancies. Obstetrics and gynecology 63:384-8, 1984
  74. Pullano JG, Cohen-Addad N, Apuzzio JJ, et al.: Water and salt conservation in the human fetus and newborn. I. Evidence for a role of fetal prolactin. The Journal of clinical endocrinology and metabolism 69:1180-6, 1989
  75. Golander A, Kopel R, Lazebnik N, et al.: Decreased prolactin secretion by decidual tissue of pre-eclampsia in vitro. Acta endocrinologica 108:111-3, 1985
  76. Healy DL, Herington AC, O'Herlihy C: Chronic polyhydramnios is a syndrome with a lactogen receptor defect in the chorion laeve. British journal of obstetrics and gynaecology 92:461-7, 1985
  77. McCoshen JA, Barc J: Prolactin bioactivity following decidual synthesis and transport by amniochorion. American journal of obstetrics and gynecology 153:217-23, 1985
  78. Rutanen E: Insulin-like growth factor binding protein-1. Semin Reprod Endocrinol 10:154, 1992
  79. Iwashita M, Kobayashi M, Matsuo A, et al.: Feto-maternal interaction of IGF-I and its binding proteins in fetal growth. Early Hum Dev 29:187-91, 1992
  80. Seppala M, Riittinen L, Kamarainen M: Placental protein 14/progesterone-associated endoemtrial protein revisited. Semin Reprod Endocrinol 10:164, 1992
  81. Julkunen M, Rutanen EM, Koskimies A, et al.: Distribution of placental protein 14 in tissues and body fluids during pregnancy. British journal of obstetrics and gynaecology 92:1145-51, 1985
  82. Stabile I, Olajide F, Chard T, et al.: Circulating levels of placental protein 14 in ectopic pregnancy. British journal of obstetrics and gynaecology 101:762-4, 1994
  83. Carr BR, MacDonald PC, Simpson ER: The role of lipoproteins in the regulation of progesterone secretion by the human corpus luteum. Fertility and sterility 38:303-11, 1982
  84. Csapo AI, Pulkkinen MO, Wiest WG: Effects of luteectomy and progesterone replacement therapy in early pregnant patients. American journal of obstetrics and gynecology 115:759-65, 1973
  85. Sauer MV, Paulson RJ, Lobo RA: A preliminary report on oocyte donation extending reproductive potential to women over 40. The New England journal of medicine 323:1157-60, 1990
  86. Yen SS: Endocrine-metabolic adaptations in pregnancy; in Reproductive endocrinology: physiology, pathophysiology and clinical management. Edited by Yen SSC, Jaffe RB, Barbieri RL. Philadelphia: WB Saunders Company, 1991
  87. Nygren KG, Johansson ED, Wide L: Evaluation of the prognosis of threatened abortion from the peripheral plasma levels of progesterone, estradiol, and human chorionic gonadotropin. American journal of obstetrics and gynecology 116:916-22, 1973
  88. Stovall TG, Ling FW, Carson SA, et al.: Serum progesterone and uterine curettage in differential diagnosis of ectopic pregnancy. Fertility and sterility 57:456-7, 1992
  89. Fields PA, Larkin LH: Purification and immunohistochemical localization of relaxin in the human term placenta. The Journal of clinical endocrinology and metabolism 52:79-85, 1981
  90. Lopez Bernal A, Bryant-Greenwood GD, Hansell DJ, et al.: Effect of relaxin on prostaglandin E production by human amnion: changes in relation to the onset of labour. British journal of obstetrics and gynaecology 94:1045-51, 1987
  91. Weiss G, O'Byrne EM, Hochman J, et al.: Distribution of relaxin in women during pregnancy. Obstetrics and gynecology 52:569-70, 1978
  92. Emmi AM, Skurnick J, Goldsmith LT, et al.: Ovarian control of pituitary hormone secretion in early human pregnancy. The Journal of clinical endocrinology and metabolism 72:1359-63, 1991
  93. Marnach ML, Ramin KD, Ramsey PS, et al.: Characterization of the relationship between joint laxity and maternal hormones in pregnancy. Obstetrics and gynecology 101:331-5, 2003
  94. Hwang JJ, Macinga D, Rorke EA: Relaxin modulates human cervical stromal cell activity. The Journal of clinical endocrinology and metabolism 81:3379-84, 1996
  95. MacLennan AH, Katz M, Creasy R: The morphologic characteristics of cervical ripening induced by the hormones relaxin and prostaglandin F2 alpha in a rabbit model. American journal of obstetrics and gynecology 152:691-6, 1985
  96. Garibay-Tupas JL, Maaskant RA, Greenwood FC, et al.: Characteristics of the binding of 32P-labelled human relaxins to the human fetal membranes. The Journal of endocrinology 145:441-8, 1995
  97. Bryant-Greenwood GD, Kern A, Yamamoto SY, et al.: Relaxin and the human fetal membranes. Reproductive sciences 14:42-5, 2007
  98. Tulchinsky D, Hobel CJ, Yeager E, et al.: Plasma estrone, estradiol, estriol, progesterone, and 17-hydroxyprogesterone in human pregnancy. I. Normal pregnancy. American journal of obstetrics and gynecology 112:1095-100, 1972
  99. Dicztalusy E: Steroid metabolism in the feto-placental unit; in The Feto-Placental Unit. Edited by Pecile A, Finzi C. Amsterdam: Excerpta Medica, 1969
  100. Pepe GJ, Albrecht ED: Actions of placental and fetal adrenal steroid hormones in primate pregnancy. Endocrine reviews 16:608-48, 1995
  101. Abraham GE, Odell WD, Swerdloff RS, et al.: Simultaneous radioimmunoassay of plasma FSH, LH, progesterone, 17-hydroxyprogesterone, and estradiol-17 beta during the menstrual cycle. The Journal of clinical endocrinology and metabolism 34:312-8, 1972
  102. Lindberg BS, Johansson ED, Nilsson BA: Plasma levels of nonconjugated oestrone, oestradiol-17beta and oestriol during uncomplicated pregnancy. Acta obstetricia et gynecologica Scandinavica 32:21-36, 1974
  103. Yen SS: Endocrine-metabolic adaptations in pregnancy; in Reproductive endocrinology: physiology, pathophysiology and clinical management. Edited by Yen SSC, Jaffe RB, Barbieri RL. Philadelphia: WB Saunders Company, 1991
  104. Mitchell BF, Challis JR, Lukash L: Progesterone synthesis by human amnion, chorion, and decidua at term. American journal of obstetrics and gynecology 157:349-53, 1987
  105. Siiteri PK, Febres F, Clemens LE, et al.: Progesterone and maintenance of pregnancy: is progesterone nature's immunosuppressant? Annals of the New York Academy of Sciences 286:384-97, 1977
  106. Moriyama I, Sugawa T: Progesterone facilitates implantation of xenogenic cultured cells in hamster uterus. Nature: New biology 236:150-2, 1972
  107. Partsch CJ, Sippell WG, MacKenzie IZ, et al.: The steroid hormonal milieu of the undisturbed human fetus and mother at 16-20 weeks gestation. The Journal of clinical endocrinology and metabolism 73:969-74, 1991
  108. Tulchinsky D, Simmer HH: Sources of plasma 17alpha-hydroxyprogesterone in human pregnancy. The Journal of clinical endocrinology and metabolism 35:799-808, 1972
  109. Siiteri PK, MacDonald PC: Placental estrogen biosynthesis during human pregnancy. The Journal of clinical endocrinology and metabolism 26:751-61, 1966
  110. Bradshaw KD, Carr BR: Placental sulfatase deficiency: maternal and fetal expression of steroid sulfatase deficiency and X-linked ichthyosis. Obstetrical & gynecological survey 41:401-13, 1986
  111. Simpson ER, Mahendroo MS, Means GD, et al.: Aromatase cytochrome P450, the enzyme responsible for estrogen biosynthesis. Endocrine reviews 15:342-55, 1994
  112. Yen SS: Endocrine-metabolic adaptations in pregnancy; in Reproductive endocrinology: physiology, pathophysiology and clinical management. Edited by Yen SSC, Jaffe RB, Barbieri RL. Philadelphia: WB Saunders Company, 1991
  113. Resnik R, Killam AP, Battaglia FC, et al.: The stimulation of uterine blood flow by various estrogens. Endocrinology 94:1192-6, 1974
  114. Henson MC, Pepe GJ, Albrecht ED: Regulation of placental low-density lipoprotein uptake in baboons by estrogen: dose-dependent effects of the anti-estrogen ethamoxytriphetol (MER-25). Biology of reproduction 45:43-8, 1991
  115. Tulchinsky D, Hobel CJ, Korenman SG: A radioligand assay for plasma unconjugated estriol in normal and abnormal pregnancies. American journal of obstetrics and gynecology 111:311-8, 1971
  116. Landon MB, Gabbe SG: Fetal surveillance in the pregnancy complicated by diabetes mellitus. Clinical obstetrics and gynecology 34:535-43, 1991
  117. Klopper A, Masson G, Campbell D, et al.: Estriol in plasma. A compartmental study. American journal of obstetrics and gynecology 117:21-6, 1973
  118. Solomon S: The placenta as an endocrine organ: steroids; in The physiology of reproduction. Edited by Knobil E, Neill JD. New York: Raven Press Ltd., 1988
  119. Barnea ER, Kaplan M: Spontaneous, gonadotropin-releasing hormone-induced, and progesterone-inhibited pulsatile secretion of human chorionic gonadotropin in the first trimester placenta in vitro. The Journal of clinical endocrinology and metabolism 69:215-7, 1989
  120. Petraglia F, Florio P, Nappi C, et al.: Peptide signaling in human placenta and membranes: autocrine, paracrine, and endocrine mechanisms. Endocrine reviews 17:156-86, 1996
  121. Chrousos GP, Calabrese JR, Avgerinos P, et al.: Corticotropin releasing factor: basic studies and clinical applications. Progress in neuro-psychopharmacology & biological psychiatry 9:349-59, 1985
  122. Stalla GK, Hartwimmer J, von Werder K, et al.: Ovine (o) and human (h) corticotrophin releasing factor (CRF) in man: CRF-stimulation and CRF-immunoreactivity. Acta endocrinologica 106:289-97, 1984
  123. Shibahara S, Morimoto Y, Furutani Y, et al.: Isolation and sequence analysis of the human corticotropin-releasing factor precursor gene. The EMBO journal 2:775-9, 1983
  124. Stalla GK, Bost H, Stalla J, et al.: Human corticotropin-releasing hormone during pregnancy. Gynecol Endocrinol 3:1-10, 1989
  125. Laatikainen TJ, Raisanen IJ, Salminen KR: Corticotropin-releasing hormone in amniotic fluid during gestation and labor and in relation to fetal lung maturation. American journal of obstetrics and gynecology 159:891-5, 1988
  126. Jones SA, Brooks AN, Challis JR: Steroids modulate corticotropin-releasing hormone production in human fetal membranes and placenta. The Journal of clinical endocrinology and metabolism 68:825-30, 1989
  127. Robinson BG, Emanuel RL, Frim DM, et al.: Glucocorticoid stimulates expression of corticotropin-releasing hormone gene in human placenta. Proceedings of the National Academy of Sciences of the United States of America 85:5244-8, 1988
  128. Linton EA, Perkins AV, Woods RJ, et al.: Corticotropin releasing hormone-binding protein (CRH-BP): plasma levels decrease during the third trimester of normal human pregnancy. The Journal of clinical endocrinology and metabolism 76:260-2, 1993
  129. Sug-Tang A, Bocking AD, Brooks AN, et al.: Effects of restricting uteroplacental blood flow on concentrations of corticotrophin-releasing hormone, adrenocorticotrophin, cortisol, and prostaglandin E2 in the sheep fetus during late pregnancy. Canadian journal of physiology and pharmacology 70:1396-402, 1992
  130. Clifton VL, Read MA, Leitch IM, et al.: Corticotropin-releasing hormone-induced vasodilatation in the human fetal placental circulation. The Journal of clinical endocrinology and metabolism 79:666-9, 1994
  131. Perkins AV, Linton EA, Eben F, et al.: Corticotrophin-releasing hormone and corticotrophin-releasing hormone binding protein in normal and pre-eclamptic human pregnancies. British journal of obstetrics and gynaecology 102:118-22, 1995
  132. Goland RS, Jozak S, Warren WB, et al.: Elevated levels of umbilical cord plasma corticotropin-releasing hormone in growth-retarded fetuses. The Journal of clinical endocrinology and metabolism 77:1174-9, 1993
  133. Ruth V, Hallman M, Laatikainen T: Corticotropin-releasing hormone and cortisol in cord plasma in relation to gestational age, labor, and fetal distress. American journal of perinatology 10:115-8, 1993
  134. Goland RS, Conwell IM, Warren WB, et al.: Placental corticotropin-releasing hormone and pituitary-adrenal function during pregnancy. Neuroendocrinology 56:742-9, 1992
  135. Youngblood WW, Humm J, Lipton MA, et al.: Thyrotropin-releasing hormone-like bioactivity in placenta: evidence for the existence of substances other than Pyroglu-His-Pro-NH2 (TRH) capable of stimulating pituitary thyrotropin release. Endocrinology 106:541-6, 1980
  136. Bajoria R, Babawale M: Ontogeny of endogenous secretion of immunoreactive-thyrotropin releasing hormone by the human placenta. The Journal of clinical endocrinology and metabolism 83:4148-55, 1998
  137. Taliadouros GS, Canfield RE, Nisula BC: Thyroid-stimulating activity of chorionic gonadotropin and luteinizing hormone. The Journal of clinical endocrinology and metabolism 47:855-60, 1978
  138. Kumasaka T, Nishi N, Yaoi Y, et al.: Demonstration of immunoreactive somatostatin-like substance in villi and decidua in early pregnancy. American journal of obstetrics and gynecology 134:39-44, 1979
  139. Tsalikian E, Foley TP, Jr., Becker DJ: Characterization of somatostatin specific binding in plasma cell membranes of human placenta. Pediatric research 18:953-7, 1984
  140. Ren SG, Braunstein GD: Human chorionic gonadotropin. Semin Reprod Endocrinol 10:95, 1992
  141. Mersol-Barg MS, Miller KF, Choi CM, et al.: Inhibin suppresses human chorionic gonadotropin secretion in term, but not first trimester, placenta. The Journal of clinical endocrinology and metabolism 71:1294-8, 1990
  142. Hanson FW, Powell JE, Stevens VC: Effects of HCG and human pituitary LH on steroid secretion and functional life of the human corpus luteum. The Journal of clinical endocrinology and metabolism 32:211-5, 1971
  143. Nisula BC, Ketelslegers JM: Thyroid-stimulating activity and chorionic gonadotropin. The Journal of clinical investigation 54:494-9, 1974
  144. Seron-Ferre M, Lawrence CC, Jaffee RB: Role of hCG in the regulation of the fetal adrenal gland. The Journal of clinical endocrinology and metabolism 46:834, 1978
  145. Huhtaniemi IT, Korenbrot CC, Jaffe RB: HCG binding and stimulation of testosterone biosynthesis in the human fetal testis. The Journal of clinical endocrinology and metabolism 44:963-7, 1977
  146. Hodgen GD, Itskovitz J: Recognition and maintenance of pregnancy; in The physiology of reproduction. Edited by Knobil E, Neill JD. New York: Raven Press Ltd., 1988
  147. Frankenne F, Closset J, Gomez F, et al.: The physiology of growth hormones (GHs) in pregnant women and partial characterization of the placental GH variant. The Journal of clinical endocrinology and metabolism 66:1171-80, 1988
  148. Eriksson L, Frankenne F, Eden S, et al.: Growth hormone secretion during termination of pregnancy. Further evidence of a placental variant. Acta Obstet Gynecol Scand 67:549-52, 1988
  149. Eriksson L: Growth hormone in human pregnancy. Maternal 24-hour serum profiles and experimental effects of continuous GH secretion. Acta obstetricia et gynecologica Scandinavica 147:1-38, 1989
  150. Eriksson L, Frankenne F, Eden S, et al.: Growth hormone 24-h serum profiles during pregnancy--lack of pulsatility for the secretion of the placental variant. British journal of obstetrics and gynaecology 96:949-53, 1989
  151. Mirlesse V, Frankenne F, Alsat E, et al.: Placental growth hormone levels in normal pregnancy and in pregnancies with intrauterine growth retardation. Pediatric research 34:439-42, 1993
  152. Takata K, Kasahara T, Kasahara M, et al.: Localization of erythrocyte/HepG2-type glucose transporter (GLUT1) in human placental villi. Cell and tissue research 267:407-12, 1992
  153. Hauguel-de Mouzon S, Leturque A, Alsat E, et al.: Developmental expression of Glut1 glucose transporter and c-fos genes in human placental cells. Placenta 15:35-46, 1994
  154. MacLeod JN, Lee AK, Liebhaber SA, et al.: Developmental control and alternative splicing of the placentally expressed transcripts from the human growth hormone gene cluster. The Journal of biological chemistry 267:14219-26, 1992
  155. de Zegher F, Vanderschueren-Lodeweyckx M, Spitz B, et al.: Perinatal growth hormone (GH) physiology: effect of GH-releasing factor on maternal and fetal secretion of pituitary and placental GH. The Journal of clinical endocrinology and metabolism 71:520-2, 1990
  156. Evain-Brion D, Alsat E, Mirlesse V, et al.: Regulation of growth hormone secretion in human trophoblastic cells in culture. Hormone research 33:256-9, 1990
  157. Han VK, Bassett N, Walton J, et al.: The expression of insulin-like growth factor (IGF) and IGF-binding protein (IGFBP) genes in the human placenta and membranes: evidence for IGF-IGFBP interactions at the feto-maternal interface. The Journal of clinical endocrinology and metabolism 81:2680-93, 1996
  158. Raben MS, Matsuzaki F, Minton PR: Growth-Promoting and Metabolic Effects of Growth Hormone. Metabolism: clinical and experimental 13:SUPPL:1102-7, 1964
  159. Salomon F, Cuneo RC, Hesp R, et al.: The effects of treatment with recombinant human growth hormone on body composition and metabolism in adults with growth hormone deficiency. The New England journal of medicine 321:1797-803, 1989
  160. Yen SS: Endocrine-metabolic adaptations in pregnancy; in Reproductive endocrinology: physiology, pathophysiology and clinical management. Edited by Yen SSC, Jaffe RB, Barbieri RL. Philadelphia: WB Saunders Company, 1991
  161. Niall HD, Hogan ML, Sauer R, et al.: Sequences of pituitary and placental lactogenic and growth hormones: evolution from a primordial peptide by gene reduplication. Proceedings of the National Academy of Sciences of the United States of America 68:866-70, 1971
  162. Braunstein GD, Rasor JL, Engvall E, et al.: Interrelationships of human chorionic gonadotropin, human placental lactogen, and pregnancy-specific beta 1-glycoprotein throughout normal human gestation. American journal of obstetrics and gynecology 138:1205-13, 1980
  163. Kim YJ, Felig P: Plasma chorionic somatomammotropin levels during starvation in midpregnancy. The Journal of clinical endocrinology and metabolism 32:864-7, 1971
  164. Furlanetto RW, Underwood LE, Van Wyk JJ, et al.: Serum immunoreactive somatomedin-C is elevated late in pregnancy. The Journal of clinical endocrinology and metabolism 47:695-8, 1978
  165. Kirwan JP, Hauguel-De Mouzon S, Lepercq J, et al.: TNF-alpha is a predictor of insulin resistance in human pregnancy. Diabetes 51:2207-13, 2002
  166. Navot D, Scott RT, Droesch K, et al.: The window of embryo transfer and the efficiency of human conception in vitro. Fertility and sterility 55:114-8, 1991
  167. Lenton EA, Neal LM, Sulaiman R: Plasma concentrations of human chorionic gonadotropin from the time of implantation until the second week of pregnancy. Fertility and sterility 37:773-8, 1982
  168. Kosasa T, Levesque L, Goldstein DP, et al.: Early detection of implantation using a radioimmunoassay specific for human chorionic gonadotropin. The Journal of clinical endocrinology and metabolism 36:622-4, 1973
  169. Cavanagh AC, Morton H: The purification of early-pregnancy factor to homogeneity from human platelets and identification as chaperonin 10. European journal of biochemistry / FEBS 222:551-60, 1994
  170. Di Trapani G, Orosco C, Perkins A, et al.: Isolation from human placental extracts of a preparation possessing 'early pregnancy factor' activity and identification of the polypeptide components. Human reproduction (Oxford, England) 6:450-7, 1991
  171. Zuo X, Su B, Wei D: Isolation and characterization of early pregnancy factor. Chinese medical sciences journal = Chung-kuo i hsueh k'o hsueh tsa chih / Chinese Academy of Medical Sciences 9:34-7, 1994
  172. Mehta AR, Eessalu TE, Aggarwal BB: Purification and characterization of early pregnancy factor from human pregnancy sera. The Journal of biological chemistry 264:2266-71, 1989
  173. Clarke FM: Identification of molecules and mechanisms involved in the 'early pregnancy factor' system. Reproduction, fertility, and development 4:423-33, 1992
  174. Chard T, Grudzinskas JG: Early pregnancy factor. Biological research in pregnancy and perinatology 8:53-6, 1987
  175. Mesrogli M, Schneider J, Maas DH: Early pregnancy factor as a marker for the earliest stages of pregnancy in infertile women. Human reproduction (Oxford, England) 3:113-5, 1988
  176. Shahani SK, Moniz CL, Bordekar AD, et al.: Early pregnancy factor as a marker for assessing embryonic viability in threatened and missed abortions. Gynecologic and obstetric investigation 37:73-6, 1994
  177. Straube W, Romer T, Zeenni L, et al.: [The early pregnancy factor (EPF) as an early marker of disorders in pregnancy]. Zentralblatt fur Gynakologie 117:32-4, 1995
  178. Hubel V, Straube W, Loh M, et al.: Human early pregnancy factor and early pregnancy associated protein before and after therapeutic abortion in comparison with beta-hCG, estradiol, progesterone and 17-hydroxyprogesterone. Experimental and clinical endocrinology 94:171-6, 1989
  179. Rees LH, Burke CW, Chard T, et al.: Possible placental origin of ACTH in normal human pregnancy. Nature 254:620-2, 1975
  180. Genazzani AR, Fraioli F, Hurlimann J, et al.: Immunoreactive ACTH and cortisol plasma levels during pregnancy. Detection and partial purification of corticotrophin-like placental hormone: the human chorionic corticotrophin (HCC). Clinical endocrinology 4:1-14, 1975
  181. Petraglia F, Sawchenko PE, Rivier J, et al.: Evidence for local stimulation of ACTH secretion by corticotropin-releasing factor in human placenta. Nature 328:717-9, 1987
  182. Nolten WE, Rueckert PA: Elevated free cortisol index in pregnancy: possible regulatory mechanisms. American journal of obstetrics and gynecology 139:492-8, 1981
  183. Prager D, Weber MM, Herman-Bonert V: Placental growth factors and releasing/inhibiting peptides. Semin Reprod Endocrinol 10:83, 1992
  184. Abe Y, Hasegawa Y, Miyamoto K, et al.: High concentrations of plasma immunoreactive inhibin during normal pregnancy in women. The Journal of clinical endocrinology and metabolism 71:133-7, 1990
  185. Tovanabutra S, Illingworth PJ, Ledger WL, et al.: The relationship between peripheral immunoactive inhibin, human chorionic gonadotrophin, oestradiol and progesterone during human pregnancy. Clinical endocrinology 38:101-7, 1993
  186. Petraglia F, Sawchenko P, Lim AT, et al.: Localization, secretion, and action of inhibin in human placenta. Science (New York, NY 237:187-9, 1987
  187. Muttukrishna S, Fowler PA, George L, et al.: Changes in peripheral serum levels of total activin A during the human menstrual cycle and pregnancy. The Journal of clinical endocrinology and metabolism 81:3328-34, 1996
  188. Bersinger NA, Smarason AK, Muttukrishna S, et al.: Women with preeclampsia have increased serum levels of pregnancy-associated plasma protein A (PAPP-A), inhibin A, activin A and soluble E-selectin. Hypertension in pregnancy 22:45-55, 2003
  189. Gagnon A, Wilson RD, Audibert F, et al.: Obstetrical complications associated with abnormal maternal serum markers analytes. Journal of obstetrics and gynaecology Canada : JOGC = Journal d'obstetrique et gynecologie du Canada : JOGC 30:918-49, 2008
  190. Petraglia F, Gallinelli A, Grande A, et al.: Local production and action of follistatin in human placenta. The Journal of clinical endocrinology and metabolism 78:205-10, 1994
  191. Mills NC, D'Ercole AJ, Underwood LE, et al.: Synthesis of somatomedin C/insulin-like growth factor I by human placenta. Molecular biology reports 11:231-6, 1986
  192. Grizzard JD, D'Ercole AJ, Wilkins JR, et al.: Affinity-labeled somatomedin-C receptors and binding proteins from the human fetus. The Journal of clinical endocrinology and metabolism 58:535-43, 1984
  193. Jonas HA, Harrison LC: The human placenta contains two distinct binding and immunoreactive species of insulin-like growth factor-I receptors. The Journal of biological chemistry 260:2288-94, 1985
  194. Napso T, Yong HEJ, Lopez-Tello J, et al.: The Role of Placental Hormones in Mediating Maternal Adaptations to Support Pregnancy and Lactation. Front Physiol 9:1091, 2018
  195. Levine RJ, Lam C, Qian C, et al.: Soluble endoglin and other circulating antiangiogenic factors in preeclampsia. The New England journal of medicine 355:992-1005, 2006
  196. Levine RJ, Maynard SE, Qian C, et al.: Circulating angiogenic factors and the risk of preeclampsia. The New England journal of medicine 350:672-83, 2004
  197. Venkatesha S, Toporsian M, Lam C, et al.: Soluble endoglin contributes to the pathogenesis of preeclampsia. Nature medicine 12:642-9, 2006
  198. Myatt L, Clifton RG, Roberts JM, et al.: First-trimester prediction of preeclampsia in nulliparous women at low risk. Obstetrics and gynecology 119:1234-42, 2012
  199. Kusanovic JP, Romero R, Chaiworapongsa T, et al.: A prospective cohort study of the value of maternal plasma concentrations of angiogenic and anti-angiogenic factors in early pregnancy and midtrimester in the identification of patients destined to develop preeclampsia. The journal of maternal-fetal & neonatal medicine : the official journal of the European Association of Perinatal Medicine, the Federation of Asia and Oceania Perinatal Societies, the International Society of Perinatal Obstet 22:1021-38, 2009
  200. Lenton EA, Grudzinskas JG, Gordon YB, et al.: Pregnancy specific beta 1 glycoprotein and chorionic gonadotrophin in early human pregnancy. Acta Obstet Gynecol Scand 60:489-92, 1981
  201. Chou JY, Plouzek CA: Pregnancy-specific 3t-glycoprotein. Semin Reprod Endocrinol 10:116, 1992
  202. Tatarinov YS: Trophoblast-specific beta1-glycoprotein as a marker for pregnancy and malignancies. Gynecologic and obstetric investigation 9:65-97, 1978
  203. Sinosich MJ, Teisner B, Folkersen J, et al.: Radioimmunoassay for pregnancy-associated plasma protein A. Clinical chemistry 28:50-3, 1982
  204. Bischof P: Pregnancy-associated plasma protein-A. Semin Reprod Endocrinol 10:127, 1992
  205. Westergaard JG, Teisner B, Sinosich MJ, et al.: Does ultrasound examination render biochemical tests obsolete in the prediction of early pregnancy failure? British journal of obstetrics and gynaecology 92:77-83, 1985
  206. Dumps P, Meisser A, Pons D, et al.: Accuracy of single measurements of pregnancy-associated plasma protein-A, human chorionic gonadotropin and progesterone in the diagnosis of early pregnancy failure. European journal of obstetrics, gynecology, and reproductive biology 100:174-80, 2002
  207. Obiekwe B, Pendlebury DJ, Gordeon YB, et al.: The radioimmunoassay of placental protein 5 and circulating levels in maternal blood in the third trimester of normal pregnancy. Clinica chimica acta; international journal of clinical chemistry 95:509-16, 1979
  208. Salem HT, Seppala M, Chard T: The effect of thrombin on serum placental protein 5 (PP5): is PP5 the naturally occurring antithrombin III of the human placenta? Placenta 2:205-9, 1981
  209. Flood JT, Hodgen GD: The physiology of fertilization, implantation and early human development; in Danforth's Obstetrics and Gynecology. Edited by Scott JR, Desaia PJ. Philadelphia: JB Lippincott Company, 1990
  210. Ahima RS, Flier JS: Leptin. Annual review of physiology 62:413-37, 2000
  211. Senaris R, Garcia-Caballero T, Casabiell X, et al.: Synthesis of leptin in human placenta. Endocrinology 138:4501-4, 1997
  212. Stock SM, Bremme KA: Elevation of plasma leptin levels during pregnancy in normal and diabetic women. Metabolism: clinical and experimental 47:840-3, 1998
  213. Tamas P, Sulyok E, Szabo I, et al.: Changes of maternal serum leptin levels during pregnancy. Gynecologic and obstetric investigation 46:169-71, 1998
  214. Tamura T, Goldenberg RL, Johnston KE, et al.: Serum leptin concentrations during pregnancy and their relationship to fetal growth. Obstetrics and gynecology 91:389-95, 1998
  215. Akerman F, Lei ZM, Rao CV: Human umbilical cord and fetal membranes co-express leptin and its receptor genes. Gynecological endocrinology : the official journal of the International Society of Gynecological Endocrinology 16:299-306, 2002
  216. Hardie L, Trayhurn P, Abramovich D, et al.: Circulating leptin in women: a longitudinal study in the menstrual cycle and during pregnancy. Clinical endocrinology 47:101-6, 1997
  217. Highman TJ, Friedman JE, Huston LP, et al.: Longitudinal changes in maternal serum leptin concentrations, body composition, and resting metabolic rate in pregnancy. American journal of obstetrics and gynecology 178:1010-5, 1998
  218. Hauguel-de Mouzon S, Lepercq J, Catalano P: The known and unknown of leptin in pregnancy. American journal of obstetrics and gynecology 194:1537-45, 2006
  219. Schanton M, Maymo JL, Perez-Perez A, et al.: Involvement of leptin in the molecular physiology of the placenta. Reproduction 155:R1-R12, 2018
  220. Chardonnens D, Cameo P, Aubert ML, et al.: Modulation of human cytotrophoblastic leptin secretion by interleukin-1alpha and 17beta-oestradiol and its effect on HCG secretion. Molecular human reproduction 5:1077-82, 1999
  221. Jansson N, Greenwood SL, Johansson BR, et al.: Leptin stimulates the activity of the system A amino acid transporter in human placental villous fragments. The Journal of clinical endocrinology and metabolism 88:1205-11, 2003
  222. Gnanapavan S, Kola B, Bustin SA, et al.: The tissue distribution of the mRNA of ghrelin and subtypes of its receptor, GHS-R, in humans. The Journal of clinical endocrinology and metabolism 87:2988, 2002
  223. Tanaka K, Minoura H, Isobe T, et al.: Ghrelin is involved in the decidualization of human endometrial stromal cells. The Journal of clinical endocrinology and metabolism 88:2335-40, 2003
  224. Fuglsang J: Ghrelin in pregnancy and lactation. Vitam Horm 77:259-84, 2008
  225. Winn VD, Haimov-Kochman R, Paquet AC, et al.: Gene expression profiling of the human maternal-fetal interface reveals dramatic changes between midgestation and term. Endocrinology 148:1059-79, 2007
  226. Jaffe RB: Neuroendocrine-metabolic regulation of pregnancy,; in Reproductive Endocrinology 4th ed Edited by Yen SSC JR, Barbieri RL (eds). Philadelphia: WB Saunders, 1999
  227. Fisher DA: Fetal and Neonatal Endocrinology; in Endocrinology. Edited by DeGroot LJ JJe. Philadelphia: WB Saunders, 2000
  228. Parker CR: The endocrinology of pregnancy; in Textbook of Reproductive Medicine. Edited by Carr BR, Blackwell RE. Norwalk: Appleton & Lange, 1993
  229. Tho SP, Layman LC, Lanclos KD, et al.: Absence of the testicular determining factor gene SRY in XX true hermaphrodites and presence of this locus in most subjects with gonadal dysgenesis caused by Y aneuploidy. American journal of obstetrics and gynecology 167:1794-802, 1992
  230. Jost A, Vigier B, Prepin J, et al.: Studies on sex differentiation in mammals. Recent progress in hormone research 29:1-41, 1973
  231. Baker TG: A Quantitative and Cytological Study of Germ Cells in Human Ovaries. Proceedings of the Royal Society of London Series B, Biological sciences 158:417-33, 1963
  232. Gondos B, Bhiraleus P, Hobel CJ: Ultrastructural observations on germ cells in human fetal ovaries. American journal of obstetrics and gynecology 110:644-52, 1971
  233. Johannison E: The foetal adrenal cortex in the human. Acta Endocrinol 58(Suppl. 130):7, 1968
  234. Honour JH, Wickramaratne K, Valman HB: Adrenal function in preterm infants. Biology of the neonate 61:214-21, 1992
  235. Hutchinson KA, DeCherney AH: The endocrinology of pregnancy; in Medicine of the mother and fetus. Edited by Reece EA, Hobbins JC. Philadelphia: JB Lippincott Company, 1999
  236. The morphological and functional development of the fetus. East Norwalk: Appleton & Lange, 1989
  237. Girard J: Control of fetal and neonatal glucose metabolism by pancreatic hormones. Bailliere's clinical endocrinology and metabolism 3:817-36, 1989
  238. MA. S: Carbohydrate metabolism: insulin and glucagon.; in Maternal Fetal Endocrinology 2nd ed Edited by Tulchinsky D LAe. Philadelpia: WB Saunders, 1994
  239. Alpert E, Drysdale JW, Isselbacher KJ, et al.: Human -fetoprotein. Isolation, characterization, and demonstration of microheterogeneity. The Journal of biological chemistry 247:3792-8, 1972
  240. Gitlin D, Perricelli A, Gitlin GM: Synthesis of -fetoprotein by liver, yolk sac, and gastrointestinal tract of the human conceptus. Cancer research 32:979-82, 1972
  241. Habib ZA: Maternal serum alpha-feto-protein: its value in antenatal diagnosis of genetic disease and in obstetrical-gynaecological care. Acta obstetricia et gynecologica Scandinavica 61:1-92, 1977
  242. Murgita RA, Tomasi TB, Jr.: Suppression of the immune response by alpha-fetoprotein on the primary and secondary antibody response. The Journal of experimental medicine 141:269-86, 1975
  243. Ferguson-Smith MA, Rawlinson HA, May HM, et al.: Avoidance of anencephalic and spina bifida births by maternal serum-alphafetoprotein screening. Lancet 1:1330-3, 1978
  244. Wald N, Cuckle H: AFP and age screening for Down syndrome. American journal of medical genetics 31:197-209, 1988
  245. Elster AD, Sanders TG, Vines FS, et al.: Size and shape of the pituitary gland during pregnancy and post partum: measurement with MR imaging. Radiology 181:531-5, 1991
  246. Leake RD, Weitzman RE, Glatz TH, et al.: Plasma oxytocin concentrations in men, nonpregnant women, and pregnant women before and during spontaneous labor. The Journal of clinical endocrinology and metabolism 53:730-3, 1981
  247. Zeeman GG, Khan-Dawood FS, Dawood MY: Oxytocin and its receptor in pregnancy and parturition: current concepts and clinical implications. Obstetrics and gynecology 89:873-83, 1997
  248. Glinoer D: The regulation of thyroid function in pregnancy: pathways of endocrine adaptation from physiology to pathology. Endocrine reviews 18:404-33, 1997
  249. Burrow GN, Fisher DA, Larsen PR: Maternal and fetal thyroid function. The New England journal of medicine 331:1072-8, 1994
  250. Lindsay JR, Nieman LK: The hypothalamic-pituitary-adrenal axis in pregnancy: challenges in disease detection and treatment. Endocrine reviews 26:775-99, 2005
  251. Rainey WE, Rehman KS, Carr BR: Fetal and maternal adrenals in human pregnancy. Obstetrics and gynecology clinics of North America 31:817-35, x, 2004
  252. Pepe GJ, Waddell BJ, Albrecht ED: The effects of adrenocorticotropin and prolactin on adrenal dehydroepiandrosterone secretion in the baboon fetus. Endocrinology 122:646-50, 1988
  253. Pepe GJ, Albrecht ED: Regulation of the primate fetal adrenal cortex. Endocrine reviews 11:151-76, 1990
  254. Albrecht ED, Pepe GJ: Placental steroid hormone biosynthesis in primate pregnancy. Endocrine reviews 11:124-50, 1990
  255. Carr BR, MacDonald PC, Simpson ER: The regulation of de novo synthesis of cholesterol in the human fetal adrenal gland by low density lipoprotein and adrenocorticotropin. Endocrinology 107:1000-6, 1980
  256. Carr BR, Porter JC, MacDonald PC, et al.: Metabolism of low density lipoprotein by human fetal adrenal tissue. Endocrinology 107:1034-40, 1980
  257. Parker CR, Jr., Carr BR, Winkel CA, et al.: Hypercholesterolemia due to elevated low density lipoprotein-cholesterol in newborns with anencephaly and adrenal atrophy. The Journal of clinical endocrinology and metabolism 57:37-43, 1983
  258. Simpson ER, Carr BR, Parker CR, Jr., et al.: The role of serum lipoproteins in steroidogenesis by the human fetal adrenal cortex. The Journal of clinical endocrinology and metabolism 49:146-8, 1979
  259. Winters AJ, Oliver C, Colston C, et al.: Plasma ACTH levels in the human fetus and neonate as related to age and parturition. The Journal of clinical endocrinology and metabolism 39:269-73, 1974
  260. Simpson ER, Parker CR, Jr., Carr BR: Role of lipoproteins in the regulation of steroidogenesis by the human fetal adrenal; in The Endocrine Physiology of Pregnancy and the Peripartal Period Vol 21 Serono Symposia Publications. Edited by Jaffe RB, Dell Acqua S. New York: Raven Press, 1985
  261. Nieman LK: The endocrinology of pregnancy; in Serono symposia in Reproductive Endocrinology
  262. Seron-Ferre M, Lawrence CC, Siiteri PK, et al.: Steroid production by definitive and fetal zones of the human fetal adrenal gland. The Journal of clinical endocrinology and metabolism 47:603-9, 1978
  263. Baird A, Kan KW, Solomon S: Role of pro-opiomelanocortin-derived peptides in the regulation of steroid production by human fetal adrenal cells in culture. The Journal of endocrinology 97:357-67, 1983
  264. Bugnon C, Lenys D, Bloch B, et al.: [Cyto-immunologic study of early cell differentiation phenomena in the human fetal anterior pituitary gland]. Comptes rendus des seances de la Societe de biologie et de ses filiales 168:460-5, 1974
  265. Katikineni M, Davies TF, Catt KJ: Regulation of adrenal and testicular prolactin receptors by adrenocorticotropin and luteinizing hormone. Endocrinology 108:2367-74, 1981
  266. Voutilainen R, Miller WL: Coordinate tropic hormone regulation of mRNAs for insulin-like growth factor II and the cholesterol side-chain-cleavage enzyme, P450scc [corrected], in human steroidogenic tissues. Proceedings of the National Academy of Sciences of the United States of America 84:1590-4, 1987
  267. Winters AJ, Colston C, MacDonald PC, et al.: Fetal plasma prolactin levels. The Journal of clinical endocrinology and metabolism 41:626-9, 1975
  268. Walsh SW, Norman RL, Novy MJ: In utero regulation of rhesus monkey fetal adrenals: effects of dexamethasone, adrenocorticotropin, thyrotropin-releasing hormone, prolactin, human chorionic gonadotropin, and alpha-melanocyte-stimulating hormone on fetal and maternal plasma steroids. Endocrinology 104:1805-13, 1979
  269. Liggins GC: Endocrinology of the foeto-maternal unit; in Human Reproductive Physiology. Edited by Sherman RP. Oxford: Blackwell Scientific Publications, 1972
  270. Baggia S, Albrecht ED, Pepe GJ: Regulation of 11 beta-hydroxysteroid dehydrogenase activity in the baboon placenta by estrogen. Endocrinology 126:2742-8, 1990
  271. Jost A: The fetal adrenal cortex; in Handbook of Physiology. Edited by Creep RO, Astwood WB. Washington, DC: Endocrinology Amer Physiol Soc, 1975
  272. Kondo S: Developmental studies on the Japanese human adrenals, I: ponderal growth. Bull Exp Biol 9:51, 1959
  273. Spector DVS: Handbook of Biological Data. WB Saunders, Philadelphia. 1956
  274. Fant M, Munro H, Moses AC: An autocrine/paracrine role for insulin-like growth factors in the regulation of human placental growth. The Journal of clinical endocrinology and metabolism 63:499-505, 1986
  275. Han VK, Lund PK, Lee DC, et al.: Expression of somatomedin/insulin-like growth factor messenger ribonucleic acids in the human fetus: identification, characterization, and tissue distribution. The Journal of clinical endocrinology and metabolism 66:422-9, 1988
  276. Sandman CA, Glynn L, Schetter CD, et al.: Elevated maternal cortisol early in pregnancy predicts third trimester levels of placental corticotropin releasing hormone (CRH): priming the placental clock. Peptides 27:1457-63, 2006
  277. Norwitz ER, Bonney EA, Snegovskikh VV, et al.: Molecular Regulation of Parturition: The Role of the Decidual Clock. Cold Spring Harb Perspect Med 5, 2015
  278. Berkowitz GS, Lapinski RH, Lockwood CJ, et al.: Corticotropin-releasing factor and its binding protein: maternal serum levels in term and preterm deliveries. American journal of obstetrics and gynecology 174:1477-83, 1996
  279. McGrath S, McLean M, Smith D, et al.: Maternal plasma corticotropin-releasing hormone trajectories vary depending on the cause of preterm delivery. American journal of obstetrics and gynecology 186:257-60, 2002
  280. Benedetto C, Petraglia F, Marozio L, et al.: Corticotropin-releasing hormone increases prostaglandin F2 alpha activity on human myometrium in vitro. American journal of obstetrics and gynecology 171:126-31, 1994
  281. Karalis K, Goodwin G, Majzoub JA: Cortisol blockade of progesterone: a possible molecular mechanism involved in the initiation of human labor. Nature medicine 2:556-60, 1996
  282. Case ML, MacDonald PC: Human parturition: distinction between the initiation of parturition and the onset of labor. Semin Reprod Endocrinol 11:272, 1993
  283. Olson DM, Zakar T: Intrauterine tissue prostaglandin synthesis: regulatory mechanisms. Semin Reprod Endocrinol 11:234, 1993
  284. Parker CR, Jr., Leveno K, Carr BR, et al.: Umbilical cord plasma levels of dehydroepiandrosterone sulfate during human gestation. The Journal of clinical endocrinology and metabolism 54:1216-20, 1982
  285. Mesiano S, Chan EC, Fitter JT, et al.: Progesterone withdrawal and estrogen activation in human parturition are coordinated by progesterone receptor A expression in the myometrium. The Journal of clinical endocrinology and metabolism 87:2924-30, 2002
  286. Csapo AI: Anti-progesterones in fertility control; in Pregnancy Termination: Procedures, Safety and New Developments. Edited by Zatuchn'i GI, Sciarra JJ, Speidel JJ. Hagerstown: Harper & Row, 1979
  287. Johnson JW, Austin KL, Jones GS, et al.: Efficacy of 17alpha-hydroxyprogesterone caproate in the prevention of premature labor. The New England journal of medicine 293:675-80, 1975
  288. Yemini M, Borenstein R, Dreazen E, et al.: Prevention of premature labor by 17 alpha-hydroxyprogesterone caproate. American journal of obstetrics and gynecology 151:574-7, 1985
  289. Meis PJ, Klebanoff M, Thom E, et al.: Prevention of recurrent preterm delivery by 17 alpha-hydroxyprogesterone caproate. The New England journal of medicine 348:2379-85, 2003
  290. Honnebier WJ, Swaab DF: The influence of anencephaly upon intrauterine growth of fetus and placenta and upon gestation length. The Journal of obstetrics and gynaecology of the British Commonwealth 80:577-88, 1973
  291. Ducsay CA, Seron-Ferre M, Germain AM, et al.: Endocrine and uterine activity rhythms in the perinatal period. Semin Reprod Endocrinol 11:285, 1993
  292. Honnebier MB, Nathanielsz PW: Primate parturition and the role of the maternal circadian system. European journal of obstetrics, gynecology, and reproductive biology 55:193-203, 1994
  293. Patrick J, Challis J, Campbell K, et al.: Circadian rhythms in maternal plasma cortisol and estriol concentrations at 30 to 31, 34 to 35, and 38 to 39 weeks' gestational age. American journal of obstetrics and gynecology 136:325-34, 1980
  294. Novy MJ: Hormonal regulation of parturition in primates; in Hormone Cell Interactions in Reproductive Tissues. Edited by Sciara J. New York: Masson Publishing, 1983
  295. Hirst JJ, Chibbar R, Mitchell BF: Role of oxytocin in the regulation of uterine activity during pregnancy and in the initiation of labor. Semin Reprod Endocrinol 11:219, 1993
  296. Haluska GJ, Novy MJ: Hormonal modulation of uterine activity during primate parturition. Semin Reprod Endocrinol 11:272, 1993

 

 

 

 

Non-Diabetic Hypoglycemia

ABSTRACT

 

Objective: To review the diagnosis, evaluation, and management of non-diabetic hypoglycemia in adults. Methods: A literature review using PubMed and Google Scholar was performed. In absence of data, clinical expert opinion was provided. Results: Hypoglycemia in an individual without diabetes is uncommon mainly because of a tightly regulated counterregulatory physiological response. A detailed medical history, review of medications and physical exam findings are critical first steps in providing guidance for further investigation in a non-diabetic person with documented hypoglycemia based on Whipple’s triad (presence of symptoms when plasma glucose concentrations are low and absence of symptoms with normalized glycemia). In this review, we highlight strategies to diagnose and treat hypoglycemic disorders in non-diabetic individuals based on underlying mechanisms. Conclusion: Evaluation and management of non-diabetic hypoglycemia should be individualized based on clinical presentation and suspected diagnoses.

 

INTRODUCTION

 

In healthy humans, glucose concentrations are efficiently maintained within a narrow range by the physiological mechanisms that respond to intermittent exogenous nutrient ingestion by enhancing glucose utilization and respond to intervals of nutrient deprivation by enhancing glucose production. Deviation of glucose from the normal range in both hyper or hypoglycemia only occurs when physiological mechanisms involved in maintaining the balance between the glucose utilization and the glucose production fail. As such, hypoglycemia is a manifestation of a heterogeneous group of underlying disorders that increase glucose utilization or reduce glucose production or a combination of both.

 

While hypoglycemia in persons without diabetes is relatively rare (1), the clinical relevance of this condition regarding patient safety, cognitive function, and quality of life is undeniable. Depending on severity and duration, hypoglycemia also can be fatal (2). Furthermore, hypoglycemia blunts defense against subsequent hypoglycemia leading to a vicious cycle of recurrent hypoglycemia (3), which in turn not only exaggerates related morbidities but also makes the diagnosis more complex.

 

Here, we review the current recommendations regarding diagnosis, pathophysiology, and management of hypoglycemia in non-diabetic individuals. Hypoglycemia in the pediatric population, hypoglycemia caused by anti-diabetic medications, and a comprehensive review of insulinomas can be found in the Endotext chapters entitled “Hypoglycemia in Neonates, Infants, and Children” (4), “Hypoglycemia During Therapy of Diabetes” (5) and “Insulinoma” (6), respectively.

 

PHYSIOLOGY / PATHOPHYSIOLOGY

 

Glucose is the main fuel for the brain since it cannot store glycogen or synthesize glucose (7,8). To minimize any disturbance in glucose supplies to the brain and cerebral function, therefore, redundant but very efficient physiological counter-regulatory responses are in place to prevent or correct hypoglycemia (8-10).

 

Prandial

 

In the prandial condition, the extent of glycemic excursion (difference between glycemic peak and nadir concentrations) is determined by the pace of food transition from the stomach into the gut (gastric emptying) as well as the net hepatic and extrahepatic glucose uptake (11). In healthy subjects, during the first 30-60 min of an oral glucose or mixed tolerance meal test, plasma glucose and insulin concentrations rise, shifting the hepatic net glucose output during fasting condition to net glucose uptake during the prandial state (12). As a result of changes in glucose kinetics in the early absorptive phase of glucose/mixed meal ingestion, plasma glucose concentrations start to decline in the latter absorptive phase falling below premeal levels in parallel with reduction in ingested glucose delivery to the gut. The regulatory mechanisms that are responsible for preventing hypoglycemia and restoring euglycemia during the transition of glucose flux from the ingested glucose delivery to the hepatic glucose production are not fully characterized, but both hormonal and non-hormonal factors play a role (13).

 

Fasting

 

In the fasting or postabsorptive state (4-6 hours after nutrient ingestion), plasma glucose  concentrations range from 80-90 mg/dl and rates of glucose utilization and production are equal (2mg/kg/min) (14). Glucose homeostasis during fasting is tightly regulated by a reciprocal bihormonal response, in which reduction in glucose concentrations below baseline reduces β-cell insulin secretion and stimulates α-cell glucagon release (15). While the full range of glucoregulatory effects of glucagon in the prandial state is unclear, the main function of glucagon in the fasting state is to counterbalance the action of insulin on hepatic glucose production (15-17). Glucose production is mainly (~80%) attributed to hepatic glycogenolysis with a smaller contribution (~20%) from hepatic gluconeogenesis primarily from amino acids and lactate (11). After an overnight fast, the liver contains ~ 50 g of glycogen storage, which can supply glucose for 24 hours after complete depletion. With prolonged fasting, hyperglucagonemia enhances gluconeogenesis and hypoinsulinemia promotes lipolysis.  Lipolysis releases glycerol, a gluconeogenic substrate, and free fatty acids that are converted to ketones, mainly beta hydroxybutyrate (BOHB) and acetoacetate, in the liver to be used as an alternate fuel by the brain.

 

Using a hyperinsulinemic hypoglycemic clamp in the fasting condition in normal humans, a decrement in plasma glucose concentration from the physiological range, in a hierarchical manner: (1) suppresses endogenous insulin secretion to lower glucose utilization, (2) increases glucagon response to increase glucose production, and (3) enhances epinephrine secretion (more relevant in absence of glucagon secretion) as well as cortisol and growth hormone release (more relevant during prolonged hypoglycemia) (8). Under physiological conditions during the fasting state, this counterregulatory response can reestablish euglycemia and prevent symptoms. A greater decline in plasma glucose concentration, though, would result in a symptomatic autonomic response to warn of low glucose and prompt the person to correct it by eating. However, if the individual did not (or could not) intervene, such as in the presence of overwhelming hyperinsulinemia) or blunted glucose counterregulatory responses, plasma glucose concentrations would drop further and neuroglycopenic symptoms and cognitive dysfunction would occur (Table 1) (18).

 

Table 1. Symptoms of Hypoglycemia

Autonomic (neurogenic)

Neuroglycopenic

Sweating

Anxiety

Tremor

Palpitation

Hunger

Tingling

Ill-defined symptoms

Warmth

Behavioral changes

Blurred vision

Confusion/difficulty speaking

Dizziness/lightheadedness

Lethargy and weakness

Seizure

Loss of consciousness/coma

 

It is unclear whether this hierarchy in hormonal responses or glycemic thresholds described during the fasting state would also apply to the prandial state.  Here, we classify the hypoglycemic disorders based on timing from meal ingestion (fasting versus prandial) given the differences in the regulatory factors involved in glucose metabolism between the two conditions.

 

DIAGNOSIS AND EVALUATION

 

Diagnosis of hypoglycemia should be made when symptoms accompany low plasma glucose concentrations (chemical hypoglycemia) but symptoms are absent when plasma glucose levels are normalized (Whipple’s triad (19)).

 

Chemical hypoglycemia has been defined based on a glucose threshold that can evoke a counterregulatory response. In the fasting state, a decline in glucose below 55 mg/dl causes neurogenic symptoms while insulin secretion is maximally suppressed (insulin < 3 uU/ml and C-peptide < 0.6 ng/ml) and glucagon response is maximized (9).

 

Awareness of hypoglycemia alerting individuals to correct hypoglycemia is mainly because of increased autonomic nervous system activity triggered by hypoglycemia. The autonomic (neurogenic) hypoglycemic symptoms can be adrenergic (such as palpitation, tremor, and anxiety) or cholinergic (such as sweating, hunger, and paresthesia) (18). As glucose concentrations drop below 48-50 mg/dl during the fasting state, the neuroglycopenic symptoms (caused by brain glucose deprivation) manifest; these symptoms range from behavioral changes, fatigue, and confusion to loss of consciousness or seizure (18,20-22).

 

Diagnosis of hypoglycemia in the prandial state is much more complex since the glycemic threshold to define hypoglycemia in the prandial state has not been well characterized. Using an oral glucose challenge in 650 healthy individuals in a previous study (23), 10% of subjects developed postprandial nadir glucose concentrations below 47 mg/dl without associated symptoms, suggesting that asymptomatic low glucose events are relatively common following an oral glucose load in normal humans. Hence, the mixed meal test (described below) is the preferred provocative test to diagnose prandial hypoglycemia. Furthermore, recurrent postprandial symptoms suggestive of hypoglycemia but not associated with low glucose concentrations have also been observed in normal individuals (24), indicating that other factors beyond hypoglycemia play a role in provoking autonomic symptoms.

 

Therefore, after obtaining a detailed medical history and physical exam, diagnosis of hypoglycemia should be confirmed by verification of low glucose concentration associated with symptoms or signs that are relieved by raising glucose values (Whipple’s triad).

 

A careful history of nutritional status, current medication use, and concurrent multisystem illnesses such as liver, heart, kidney failure, or sepsis, as well as a thorough physical exam and laboratory data, can point to existing primary conditions that predispose to hypoglycemia. This is especially crucial in patients who are often too ill to be subjected to extensive evaluation. For healthy subjects who lack any background predisposing illnesses, the details about timing (relationship to food ingestion, physical activities, day versus nocturnal time), severity (frequency, presence of neuroglycopenia, and requiring assistance to treat), and time of onset of hypoglycemic episodes are critical in differential diagnosis. To understand the pattern of hypoglycemic episodes, reviewing the records of symptoms, activity, food intake along with capillary blood or interstitial glucose levels measured by glucometer or continuous glucose monitoring (CGM), respectively, may be helpful. However, the accuracy of glucometer and CGM is low in the hypoglycemic range, and they should not be used for diagnostic purposes.  On the other hand, masked (blinded) monitoring by CGM can provide insights into patterns of hypoglycemic episodes and triggering factors during patients’ daily routine (25).

 

The flowchart in Figure 1 demonstrates the suggested approach for evaluation in healthy appearing patients after a careful medical history, physical exam, and laboratory data excludes an underlying illness that can predispose to hypoglycemia. Disorders that may cause hypoglycemia are listed in Table 2.

 

To confirm the diagnosis and explore etiology, it is necessary to collect blood samples during hypoglycemia, whether it occurs spontaneously or by provoked testing that can be selected based on clues from the medical history.

 

In asymptomatic patients with documented chemical hypoglycemia, artifactual hypoglycemia due to conditions such as reticulocytosis (polycythemia, sickle cell anemia), leukocytosis (leukemia), and thrombocytosis that increase in vitro glycolysis in the blood sample while awaiting laboratory analysis should be considered (26).  Also, nadir glucose levels in the prandial state can be low without any associated symptoms, particularly in persons with a history of upper gastrointestinal (GI) surgery (25). A potential diagnostic challenge in using clinical criteria remains in patients who are adapted to recurrent hypoglycemia by blunted autonomic response, so called hypoglycemic unawareness (27).  It has been well recognized that antecedent insulin-induced hypoglycemia impairs counterregulatory glucose responses and blunts hypoglycemia symptoms (mainly autonomic symptoms) in normal humans (28). Therefore, in patients with a high index of clinical suspicion, monitoring of symptoms and signs of neuroglycopenia, which is less likely to be affected by recurrent hypoglycemia, and reevaluation over time should be considered.

 

Figure 1. Evaluation of non-diabetic hypoglycemia in healthy appearing adults.

 

 

Table 2. Causes of Hypoglycemia

Artifactual Hypoglycemia (without symptoms)

Reticulocytosis (polycythemia, sickle cell anemia)

Leukocytosis (leukemia)

Thrombocytosis

Fasting Hypoglycemia (> 5 hour from the last meal)

High Insulin, Low beta-hydroxy butyrate, High glucagon response

Insulinoma

Auto immune syndrome (antibodies to insulin or the insulin receptor)

Factitial due to exogenous insulin 

Factitial due to insulin secretagogues

Induced by non-diabetic medications

Low Insulin, High beta-hydroxy butyrate, Low glucagon response

Ketotic hypoglycemia

Prolonged exercise

Alcohol induced

Glycogen storage diseases

Post Prandial Hypoglycemia (within 5 hours from the last meal)

Bariatric surgery

Nesidioblastosis

Hereditary fructose intolerance

Associated with Other Disorder

Critical illness (liver failure, congestive heart failure, sepsis, renal failure, etc.)

Malnutrition

Adrenal insufficiency

Non-islet cell tumors

 

Fasting Hypoglycemia

 

In patients with concern for fasting hypoglycemia, confirmation of Whipple’s triad and exploration of the cause is recommended during an episode of spontaneous hypoglycemia or with a supervised fast of up to 72 hours (9)(Figure.1).  During the fast, patients can consume non-caloric caffeine-free beverages with all non-essential medications discontinued.  Plasma glucose, insulin, c-peptide, and BOHB are collected every 6 hours until plasma glucose is < 60 mg/dL; at that time, frequency of blood collection should be increased to every 1-2 hours. The fast is terminated after collecting the last blood sample when the plasma glucose is < 45 mg/dL and the patient has signs and/or symptoms of hypoglycemia or if the patient has not exhibited symptoms after 72 hours have elapsed.  Alternatively, the fast can be terminated when plasma glucose  is < 55 mg/dL in men and < 35 mg/dl in women, given the sex differences in abnormal fasting glycemic concentrations (29), without signs/symptoms if Whipple’s triad was documented previously but blood samples were not collected (9). At the end of the fast, glucose response to 1 mg of glucagon IV bolus injection will be measured every 10 minutes for a 30-minute period and then the patient is fed. Insulin antibodies from baseline blood samples as well as hypoglycemic anti-diabetic medications (sulfonylureas and meglitinides) screening from baseline blood and urine samples are also collected. After confirmation of diagnosis, the results of the fasting test will help to differentiate hypoglycemia mediated by insulin- versus non-insulin factors.  

 

INSULIN-DEPENDENT HYPOGLYCEMIA (HIGH PLASMA INSULIN CONCENTRATION)

 

C-peptide is secreted from β-cells at an equimolar ratio to insulin (30). Approximately half of the insulin which is secreted into the portal vein is removed by the liver (31). Therefore, plasma insulin concentration reflects not only insulin secretion or exogenous insulin administration, but also hepatic insulin degradation. In contrast to insulin, c-peptide undergoes minimal extraction by the liver and other organs (32), therefore c-peptide concentration represents endogenous insulin secretion (32,33). It has been well documented that exogenous insulin administration during euglycemia or hypoglycemia inhibits endogenous insulin secretion (34,35). In fact, this physiological phenomenon has been used to support the diagnosis of insulinoma using hyperinsulinemic hypoglycemic or euglycemic clamp (36,37).

 

Therefore, low plasma concentration of c-peptide in presence of elevated insulin values during hypoglycemia indicates factitial hypoglycemia due to exogenous insulin administration, whereas elevated c-peptide and insulin represents inappropriately greater endogenous insulin secretion due to insulin secretagogues (sulfonylurea or meglitinides), autoimmune syndromes (insulin antibody syndrome or type B insulin resistance) or an insulin-producing tumor (insulinoma).

 

Further, hypoglycemia induced by hyperinsulinemia is associated with low BOHB and glycemic response >25 mg/dl to glucagon injection. Insulin increases fatty acid synthesis and esterification and decreases fatty acid oxidation and ketogenesis in the liver, leading to lower plasma concentrations of BOHB. Enhanced insulin signaling in the liver activated by hyperinsulinemia or any non-insulin ligands, such as insulin-like growth factor—2 (IGF-2), also results in higher glycogen storage, hence a larger glycemic response to glucagon injection.

 

Factitial Hypoglycemia

 

Factitial hypoglycemia due to exogenous insulin (high insulin and low c-peptide) or insulin secretagogues medications (high insulin and c-peptide) remains a diagnostic challenge and often leads to extensive and costly investigation to rule out other causes because of limitations in biochemical assays (38), as well as patients’ denial of medication misuse. Factitial hypoglycemia is observed more often in patients who work in the medical health care system, have relatives with diabetes living in the same household, and those with underlying mental illness such as major depression. Therefore, obtaining detailed information regarding the patient’s medication list including herbal preparations that can be contaminated with sulfonylurea as well as family history is essential (39).

 

Hypoglycemia due to exogenous insulin is characterized by elevated plasma insulin, suppressed c-peptide and low BOHB, as well as an increase in plasma glucose > 25mg/dL after glucagon challenge (9).  Insulin antibodies may also be positive (40). 

 

Insulin-induced hypoglycemia due to insulin secretagogues (sulfonylurea or meglitinide) has a similar biochemical profile except that the c-peptide is elevated.  Therefore, the only way to differentiate anti-diabetic factitial hypoglycemia from insulinoma is by detecting the drug in blood or urine.

 

In a single-center retrospective study, factitious hypoglycemia accounted for 11 of 70 (16%) of admissions for evaluation of hypoglycemia. (41).  Prognosis is poor based on a small-size study, in which only 30% of affected patients during several years of follow-up recovered (42). Treatment requires a multisystemic treatment team led by a psychiatrist (42).

 

Autoimmune Syndromes

 

Autoimmune syndromes are a rare cause of hypoglycemia characterized by high concentrations of insulin autoantibodies (insulin autoimmune syndrome [IAS]) or anti-insulin receptor antibodies (type B insulin resistance) (43,44). While IAS is the third leading cause of hypoglycemia in Japan, it is very uncommon in the non-Asian population; type B insulin resistance is even less common (43). 

 

Insulin autoantibodies (IAAs) are mainly immunoglobulins (Ig) directed against endogenously released insulin in response to nutrient ingestion with a high binding capacity but low affinity to insulin. As a result, patients may manifest hyperglycemia in the early absorptive phase of meal or oral glucose intake, when exogenous glucose appearance into circulation is maximal, followed by hypoglycemia in a few hours during the late prandial condition or postabsorptive state. Hypoglycemia, in IAS, is caused by the binding and release of insulin from the antigen-antibody complex independently of changes in glucose concentrations. Therefore, insulin and c-peptide are both elevated at the time of low plasma glucose concentrations.  IAAs are different from insulin antibodies produced against exogenous insulin that are generally low binding and high affinity, thus, unable to cause hypoglycemia.

 

In non-Asian patients, IAS is mainly reported in individuals with autoimmune (lupus, rheumatoid arthritis) /hematological diseases (multiple myeloma, benign monoclonal gammopathy), who are exposed to triggering factors, such as medications (captopril, propylthiouracil, penicillin G) and viral infections (measles, mumps, rubella, varicella zoster, coxsackie B, and hepatitis C) (43,44).  In Japanese patients, IAS is commonly associated with exposure to medications with a sulphydryl group (methimazole) (43). The prognosis is relatively good with self-remission reported in 82% of patients (44). Treatment is often dietary modification (small, frequent low-carbohydrate meals and uncooked cornstarch) and occasionally requires medications to decrease insulin secretion (somatostatin analogues, diazoxide) or immunosuppressants (high -dose corticosteroids, azathioprine, rituximab) (43,44).

 

Type B insulin resistance is caused by anti-insulin receptor antibodies. Affected patients tend to be middle-aged women of Black race with obesity, acanthosis nigricans, and hyperandrogenism.  Co-occurrence of systemic autoimmune disease (i.e., lupus) is common (43). Patients typically present with hyperglycemia; however, a subset of patients (8 of 34 patients in an NIH cohort) experience fasting or postprandial hypoglycemia after period of hyperglycemia or without a history of hyperglycemia (43). The autoantibodies are believed to be partial agonists for the insulin receptor. Hyperglycemia or hypoglycemia ensues depending on the antibody titer: high titers antagonize the receptor, resulting in hyperglycemia and high insulin and c-peptide levels to compensate for the resistance; low titers activate the receptor, leading to hypoglycemia. There are diagnostic challenges as immunoprecipitation, the gold standard method to detect insulin receptor autoantibodies, is generally not commercially available (45).  Unlike IAS, prognosis of type B insulin resistance is poor with high mortality especially in patients that transition from a hyperglycemic to hypoglycemic phase.  Deaths are related to hypoglycemia and other causes (lupus, renal failure, cancer, cardiovascular events).  Therapy can include immunosuppressants, but response is variable or poor (43).

 

Insulinoma

 

Neuroendocrine insulin-producing tumors (insulinoma) are relatively rare with an estimated incidence of ~ 1 – 4 new cases per million people/year (46). Less than 10% of insulinomas are malignant, 10% multiple, and 4% associated with multiple endocrine neoplasia type 1 (MEN-1) syndrome (47).  They primarily manifest in the 5th decade of life and are slightly more common in females (48).  Typically, patients experience episodes of hypoglycemia fasting or after exercise, but some individuals may experience fasting and prandial hypoglycemia (49).  Less frequently (6% of 237 patients in a Mayo Clinic cohort), patients present solely with prandial hypoglycemia (50).  Diagnosis relies on biochemical testing. The 72-hour fast is usually successful in capturing hypoglycemia in patients with insulinoma, as 65% of patients will experience hypoglycemia within 24 hours, 93% within 48 hours, and 99% within 72-hours (51).

 

Individuals with confirmed fasting (or postprandial) hyperinsulinemic hypoglycemia, negative screening for oral hypoglycemic medications, and negative insulin autoantibody testing should undergo diagnostic tests to locate the insulinoma prior to surgery (9).  Non-invasive imaging can include transabdominal and endoscopic ultrasonography, abdominal computed tomography (CT), and magnetic resonance imaging (MRI) (52). CT has been shown to detect ~70-80% of tumors and MRI 85% (47). Given that insulinomas tend to be small in size (< 1 cm in diameter in 40% of diagnosed cases) (47), negative imaging does not rule out the diagnosis (9). If non-invasive imaging cannot determine preoperative localization, selective arterial calcium stimulation with hepatic venous sampling can been utilized (53). Calcium is injected into arteries supplying the pancreas, which stimulates insulin secretion from insulinomas; a ≥ 2-fold increase in insulin concentrations from baseline localizes the site of an insulinoma with a > 90% sensitivity (54,55).

 

Treatment is surgical enucleation of the tumor, performed more commonly via open surgery than laparoscopic (48). If not localized prior to surgery, intraoperative palpation by an experienced surgeon coupled with intraoperative ultrasound detects > 80% of tumors (48). Pancreatectomy (distal or central depending on location of tumor) is preferred over enucleation if the insulinoma is large or there is concern for malignancy or metastases. Pancreatic fistula is the most common complication in both open and laparoscopic surgeries.  As noted earlier, the hyperinsulinemic hypoglycemic or euglycemic clamp has also be utilized to differentiate insulinoma from other causes when conventional evaluation (prolonged fasting, imaging) were equivocal (37) or when surgery failed to detect an insulinoma but clinical suspicion was high (56).

 

Benign insulinomas have a high 5yr survival rate of 95-100% post-resection with relapse frequency 6% at 10yr and 8% at 20 yrs (48). Malignant insulinomas carry a poor prognosis with median life expectancy of 2 years (48). Inoperable cases may benefit from medical management such as diazoxide, streptozocin, verapamil, and phenytoin by reducing insulin secretion and corticosteroids by diminishing insulin action (48,57-59). Non-surgical procedures such as CT guided radiofrequency ablation, US-guided ablation with ethanol, peptide receptor radionuclide therapy, and robotic radiosurgery have also been utilized (48).

 

Non-Diabetic Medications

 

Non-diabetic medications such as quinolones, non-steroidal anti-inflammatory drugs, antipsychotics, and α and β blockers have been implicated in inducing hypoglycemia by stimulating insulin secretion (60).  Other medications, such as pentamidine, damage the β cells, resulting in transient hypoglycemia induced by cytolytic insulin release (60).  Risk factors related to hypoglycemia include higher medication doses, concomitant renal failure, older age, and poor nutrition  (60).  Treatment may include discontinuation of the medication and supportive care (60).  

 

INSULIN-INDEPENDENT HYPOGLYCEMIA (LOW PLASMA INSULIN CONCENTRATION)

 

In this group, plasma BOHB is elevated and glucose response to glucagon is small. In patients who appear healthy the following conditions should be considered and ruled out: 

 

Ketotic Hypoglycemia

 

Ketotic hypoglycemia is a relatively rare condition that can occur during extended periods of carbohydrate deprivation, as during fasting or starvation. Prolonged adherence to a ketogenic diet, which severely restricts carbohydrate intake to 20-50g/day to promote weight loss and increase plasma ketone bodies (61), combined with other factors interfering with counterregulatory response, such as alcohol intake, also can result in ketotic hypoglycemia (62). Hypoglycemia in this condition is mainly caused by reduction in hepatic glucose production by hyperketonemia, which outweighs the diminished glucose utilization (63).  Management is largely supportive depending on the severity of the presentation.

 

Prolonged Exercise

 

Prolonged exercise can result in plasma glucose to decline to the hypoglycemic range in 30-40% of healthy subjects, but these events are not associated with symptoms (Whipple’s triad) nor incompatible with continued exercise (64). While the mechanisms for lack of symptoms is not completely understood, studies in non-diabetic dogs using a hypoglycemic clamp with and without exercise have shown that (a) counterregulatory hormonal responses to exercise and hypoglycemia combined are greater than the response to either conditions alone, and (b) larger insulin action during exercise are negated by the counterregulatory response during hypoglycemia (both enhanced endogenous glucose production and reduced glucose utilization) (65). Thus, disturbed counterregulatory response in prolonged exercise combined with inadequate carbohydrate ingestion may result in clinical hypoglycemia (66).  Management is supportive.

 

Alcohol-Induced Hypoglycemia

 

Hypoglycemia due to alcohol has been attributed to inhibition of gluconeogenesis and blunting of growth hormone response to hypoglycemia (67). Management is mainly supportive depending on the severity and length of hypoglycemia. 

 

Glycogen Storage Diseases

 

Glycogen storage diseases (GSD) are rare genetic disorders that impair the breakdown of glycogen.  Although these diseases are commonly diagnosed in infancy, GSDs Type I (deficiency in glucose-6-phosphatase), III (deficiency in amylo-1,6-glucosidase), and 0 (lack of glycogen synthase) can present in adulthood or continue to persist in adulthood (68).  Adults with Type 1 GSD (most common form, annual incidence 1/100,000) may present with hypoglycemia, lactic acidosis, hyperuricemia, hypertriglyceridemia, and hepatomegaly (68).  Patients with glycogen storage disease are generally managed by frequent feeding with complex carbohydrates and cornstarch to prevent hypoglycemia.  Patients need to be followed long-term by a metabolic specialist.

 

Fatty Acid Oxidation (FAO) Disorders

 

FAO disorders are a rare group of autosomal recessive conditions characterized by impaired breakdown of fatty acids, leading to hypoketotic hypoglycemia and myopathy. FAO disorders also typically manifest in childhood but can continue through adulthood. Prognosis depends on the specific condition and severity. Treatment typically includes avoidance of fasting and high carbohydrate/low fat diets (68). Patients are managed by metabolic specialists. 

 

ASSOCIATED WITH OTHER DISORDERS

 

In ill-appearing patients with hypoglycemia but low insulin concentrations the following conditions should be considered:

 

Critical Illness

 

Critical illness including organ failure such as acute liver failure and congestive heart failure with hepatic congestion have been associated with hypoglycemia, likely due to impaired gluconeogenesis and depletion of hepatic glycogen stores (69,70).  Sepsis-induced hypoglycemia has been appreciated in humans and animal models with depleted glycogen stores, impaired gluconeogenesis, and increased peripheral glucose utilization implicated as contributing factors (71,72).  Hypoglycemia in non-diabetic people with end stage renal disease is attributed to concomitant adrenal insufficiency, certain medications, malnutrition, and infection (73,74).  Management of Ill-appearing individuals due to sepsis or organ failure is mainly treatment of underlying disorders and treatment of severe hypoglycemia by intravenous glucose administration.

 

Addison’s Disease

 

Addison’s disease is a rare disease that results in primary adrenal insufficiency characterized by glucocorticoid deficiency with or without mineralocorticoid deficiency.  Most cases are caused by autoimmune damage to the adrenal cortex by 21-hydroxylase antibodies (75).  Hypoglycemia, although rare, is likely due to cortisol deficiency that interferes with counterregulatory response during times of stress (76).  Addison’s disease should be considered in a hypoglycemic individual with hyperpigmentation, hyponatremia, hyperkalemia and acidosis (75). Further testing, such as ACTH stimulation test, may be warranted to confirm diagnosis (77).  Treatment for patients with primary adrenal insufficiency in adrenal crisis should be initiated by volume replacement and immediate treatment with intravenous or intramuscular hydrocortisone 100 mg followed by 100 mg every 6-8 hours until clinically stable.  Etiology of the precipitating adrenal crisis should be identified and treated (i.e., infection, hemorrhage, etc.) (75).

 

Non-Islet Cell Tumors

 

Nonislet cell tumors are rare mesenchymal and epithelial tumors that can be benign or malignant, are often large (> 10 cm), and clinically apparent.  Hypoglycemia results when the tumor overproduces incompletely processed IGF-2 (9,78-80). Tumors can secrete IGF-2 or its posttranslational precursor “big IGF” (81).  IGF-2 structurally is similar to insulin; high levels of IGF-2 can bind to the insulin receptor and mimic the action of insulin, resulting in hypoglycemia (82). Insulin and c-peptide levels are appropriately suppressed in response to the hypoglycemia.  Management of non-islet cell tumors may include surgery, radiotherapy, chemotherapy and medical therapy with glucocorticoids, GH, or octreotide (9).

 

Prandial Hypoglycemia

 

The glycemic threshold to define hypoglycemia after meal ingestion is unknown. However, symptoms associated with plasma glucose less than 50-55 mg/dL during mixed meal test that is relieved by normalization of glucose has been used to confirm meal-induced hypoglycemia (11,25).  In the prandial state, provocative testing should use a mixed meal containing protein, carbohydrates, and fat and not oral glucose. This is mainly because the oral glucose challenge has low specificity for detecting clinical hypoglycemia by causing asymptomatic low glucose nadirs as well as hypoglycemia symptoms (mainly autonomic) that do not correlate with low glucose concentrations (23,24).

 

Currently meal tests are not standardized as both solid and liquid mixed meals as well as variable carbohydrate content from 45 to 105 grams have been used (83). Regardless of approach, meal studies can increase the risk of inducing hypoglycemia, thus, these tests need to be done under supervision by personnel trained in a safe environment.

 

HYPOGLYCEMIA AFTER BARIATRIC SURGERY

 

Meal-induced hypoglycemia after upper GI tract (gastrectomy and pyloroplasty) (84) or bariatric surgery (85-87) (Roux-en y gastric bypass surgery [RYGB] and sleeve gastrectomy [SG]) are well documented. One in 10 bariatric subjects develop a late-complication of hypoglycemia (88,89), and one in 150 suffer from severe hypoglycemia requiring an emergency room visit or hospitalization (88). Hypoglycemia in this population is postprandial, progressive, often associated with cognitive impairment and occasionally with loss of consciousness or seizures and is only partially responsive to diet modification or available therapeutic options (25,86,90-92). Despite sporadic case reports of postprandial hypoglycemia after SG (85), this condition, in our experience, is less prevalent, and likely to be of lesser severity than RYGB. Severe hypoglycemia after bariatric surgery is debilitating as it compromises patient safety, cognition, and quality of life (both professional and personal). The long-term health outcomes of this debilitating complication are largely unknown. A recent study (93) using a driver simulator has demonstrated that driving performance and cognitive function is impaired following RYGB during prandial hypoglycemia without any changes in perception of symptoms (94,95).

 

Differentiating true hypoglycemia from those with prandial asymptomatic low glucose concentration or prandial symptoms without low glucose levels is more challenging in patients after GI surgery than non-operated individuals because of higher frequency of both conditions after bariatric surgery. Using CGM for 5 days has demonstrated that 70% of non-diabetic subjects after RYGB (n=40) had at least one episode of low interstitial glucose concentration (<55 mg/dl) (94). However, 80% of these low glucose events have been shown to be asymptomatic. Furthermore, it is well documented that a large proportion of bariatric patients experience dumping symptoms (91), which are almost identical with autonomic symptoms of hypoglycemia, but not associated with low glucose concentrations. Therefore, it is critical to document Whipple’s triad (neuroglycopenic rather than autonomic symptoms associated with low glucose) during free-living conditions or using mixed meal test to confirm hypoglycemia in this population.

 

Additional testing should be considered in ill-appearing patients after bariatric surgery to exclude adrenal insufficiency, other critical illnesses, and malnutrition (25). Post bariatric patients who experience fasting hypoglycemia (beyond 5 hours from previous meal ingestion) or hypoglycemia within 6-12 months from surgery should be evaluated for other causes of hypoglycemia such as insulinoma (25).

 

Underlying mechanisms by which rerouted gut after GI surgeries cause hypoglycemia is not completely understood. However, it is well documented that following RYGB, and to a smaller extent after SG, meal ingestion enhances glucose excursion leading to higher glucose peaks and lower nadir glucose concentrations mainly due to faster nutrient emptying from the stomach pouch/stomach to the gut (96) (Figure 2). Increased glucose delivery from the stomach pouch/ tube-like stomach to the gut after bariatric surgery is associated with hyperinsulinemia, which is exaggerated in RYGB patients with hypoglycemia compared to asymptomatic RYGB subjects (92) (Figure 2). Enhanced meal-induced beta-cell secretion in patients with hypoglycemia after RYGB has been attributed to not only a greater beta-cell sensitivity to increasing glycemia in the first absorptive phase, but also a lower insulin suppression during glycemic decline from peak to glucose nadir (87).

 

Prandial hyperinsulinemia after RYGB, particularly in patients with hypoglycemia, has been shown to be associated with greater prandial plasma concentration of glucagon-like peptide 1 (GLP-1), an insulinotropic gut hormone (87,90,97)(Figure 2). These observations hinted towards a key role for GLP-1 signaling beyond glycemic stimuli in meal-stimulated hyperinsulinemia after RYGB. In fact, we and others have shown that blocking the GLP-1 receptor (GLP-1R) corrects post-RYGB hypoglycemia (87,98).

 

Figure 2. The prandial glycemic effects of RYGB are exaggerated in patients with late-complication of hypoglycemia. RYGB enhances prandial glycemic excursion and increases insulin secretion rate (ISR) along with plasma GLP-1 concentrations. Patients with documented hypoglycemia after RYGB have greater insulin and GLP-1 secretin compared to those without. Following RYGB, glucagon response to meal ingestion is enhanced but there is no further increase in response to hypoglycemia. Adapted with permission from Salehi, JCEM, 2014.

 

Despite a larger meal-induced glucagon response after RYGB compared to non-operated individuals, there is no further increase in plasma glucagon concentration during prandial hyperglycemia (Figure 2), suggestive of dysregulated pancreatic α-cell response. In fact, we have shown that patients with RYGB and SG, glucagon response to insulin-induced hypoglycemia is smaller than healthy individuals without GI surgery (99,100). This data is aligned with a report demonstrating that counterregulatory hormonal response (glucagon, cortisol, and catecholamines) to hypoglycemia is significantly reduced after RYGB compared with before surgery (101). In prandial state using tracer technique, we also have shown that despite a larger prandial plasma glucagon concentration, endogenous glucose production response to hypoglycemia is smaller after RYGB compared to non-operated controls (102), suggestive of diminished liver sensitivity to glucagon. However, we have demonstrated that blocking the GLP-1R increases the prandial hepatic glucose production response to insulin-induced hypoglycemia in RYGB subjects but not in non-operated controls, suggesting that enhanced GLP-1 signal due to rerouted gut can potentially contribute to the impaired counterregulatory response to hypoglycemia (103).

 

Based on current pathophysiologic understanding treatment strategies that selectively reduce the pace of nutrient delivery to the gut and prandial insulin secretion or improve counterregulatory response are the most effective options. Because of limited therapeutic options at this point, dietary modification remains the cornerstone of management. The goal of dietary modification is to lower prandial glucose spikes while increasing glucose nadirs by (a) lowering the amount of carbohydrates for every meal (<30g) or snacks (<15 g), (b) avoiding simple carbohydrates with high-glycemic index, (c) adding fats and proteins to every meal and snack, and (d) changing the composition of carbohydrate from glucose to fructose (25). Uncooked starch has also been used in this population based on effectiveness in reducing hypoglycemic episodes in patients with diabetes mellitus (104).

 

The current medical interventions rely on drugs that had previously been used for treatment of other hypoglycemic conditions.  Acarbose, an antidiabetic medication, has been utilized as the first drug started with dietary modification. The effect of this intestinal alpha-glucosidase blocker is to block carbohydrate absorption and reduce prandial glycemic excursion after RYGB {Valderas, 2012 #6617}. Adverse effects include flatulence and bloating, especially if the dose is not gradually titrated. Other medications such as somatostatin analogues, diazoxide, and GLP-1R agonists also are used based on sporadic case reports (25). Unblinded CGM in a small size study of patients with post-RYGB hypoglycemia has also been shown to reduce hypoglycemic episodes, likely due to better self-assessment of glycemic excursion and the need for treatments (105). The investigational drugs that are in various phases of development include exendin-(9-30) (Eiger Biopharmaceutical, Paol Alto, CA, USA), a potent GLP-1R antagonist (87,106), glucagon-based drugs (107,108), somatostatin analogues (109) and sodium-glucose cotransporter-1 inhibitor (110).

 

OTHER PRANDIAL HYPOGLYCEMIC CONDITIONS (RARE)

 

Nesidioblastosis

 

Non-insulinoma pancreatogenous hypoglycemia is a rare condition that typically causes hypoglycemia in the postprandial state due to diffuse nesidioblastosis (β cell hypertrophy, islet hyperplasia, increase in β cell mass) (111). The pathogenesis in adults is largely unknown, but likely differs from nesidioblastosis that occurs in congenital hyperinsulinism that is caused by a genetic mutation (112,113). This etiology should be considered in a hypoglycemia patient with a negative 72 hour fast, positive mixed meal test concerning for endogenous hyperinsulinism without a history of GI surgery, and negative imaging for insulinoma. Selective arterial calcium stimulation demonstrates diffuse insulin secretion. Histopathology reveals nesidioblastosis (114-116). Ideal management is difficult to determine as the condition is rare; a majority of the publications are from case reports or case-series, and there is a lack of long-term follow up (82,117).  Management can include dietary interventions (low-carbohydrate frequent meals) or medical interventions with diazoxide, acarbose, verapamil or octreotide (82).  When these interventions fail, partial pancreatectomy can be performed in patients with severe neuroglycopenic symptoms (118).

 

Hereditary Fructose Intolerance

 

Hereditary fructose intolerance is a rare autosomal recessive disorder (<1-9/100,000 annual incidence) caused by fructose-1-phoshate deficiency that results in postprandial hypoglycemia after ingestion of fructose (fruits) or sucrose (sweet foods) that is usually diagnosed in childhood. The diagnosis should be suspected in ill-appearing adults with hypoglycemia associated with GI symptoms (nausea, vomiting, diffuse abdominal pain) after eating fruits or sugar. Clinical symptom resolution within days of elimination suggests hereditary fructose intolerance and can be confirmed by molecular diagnosis on DNA obtained from peripheral leukocytes (68).

 

CONCLUSION

 

Non-diabetic hypoglycemia is a rare phenomenon since in healthy individuals counterregulatory mechanisms prevent and correct hypoglycemia by reducing glucose uptake and by enhancing hepatic glucose production. These mechanisms are less well characterized in the prandial state compared to the fasting state. Nonetheless, hypoglycemia only occurs when impaired physiological responses offset the balance between glucose utilization and production. Evaluation of hypoglycemia starts with a detailed history, comprehensive review of medications and clinical presentation, and a thorough physical exam that guides the diagnostic approach in patients with documented hypoglycemia based on Whipple’s triad. In absence of confirmation of hypoglycemia during free living condition provoked testing, prolonged fast or mixed meal test for hypoglycemic conditions reported during fasting or prandial state, respectively, are indicated. Diagnostic and management strategies for non-diabetic hypoglycemia is individualized depending on specific pathophysiology and can include interventions that are dietary, medical, or surgical.

 

REFERENCES

 

  1. Cryer PE, Davis SN, Shamoon H. Hypoglycemia in diabetes. Diabetes Care. 2003;26(6):1902-1912.
  2. Cryer PE. Severe hypoglycemia predicts mortality in diabetes. Diabetes Care. 2012;35(9):1814-1816.
  3. Davis MR, Shamoon H. Counterregulatory adaptation to recurrent hypoglycemia in normal humans. J Clin Endocrinol Metab. 1991;73(5):995-1001.
  4. Rosenfeld E, Thornton PS. Hypoglycemia in Neonates, Infants, and Children. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  5. Davis HA, Spanakis EK, Cryer PE, Davis SN. Hypoglycemia During Therapy of Diabetes. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  6. de Herder WW, Hofland J. Insulinoma. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  7. Lopez-Gambero AJ, Martinez F, Salazar K, Cifuentes M, Nualart F. Brain Glucose-Sensing Mechanism and Energy Homeostasis. Mol Neurobiol. 2019;56(2):769-796.
  8. Cryer PE. Hypoglycemia, functional brain failure, and brain death. J Clin Invest. 2007;117(4):868-870.
  9. Cryer PE, Axelrod L, Grossman AB, Heller SR, Montori VM, Seaquist ER, Service FJ, Endocrine S. Evaluation and management of adult hypoglycemic disorders: an Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab. 2009;94(3):709-728.
  10. Service FJ. Hypoglycemic disorders. N Engl J Med. 1995;332(17):1144-1152.
  11. Service FJ. Hypoglycemias. West J Med. 1991;154(4):442-454.
  12. Petersen MC, Vatner DF, Shulman GI. Regulation of hepatic glucose metabolism in health and disease. Nat Rev Endocrinol. 2017;13(10):572-587.
  13. Tse TF, Clutter WE, Shah SD, Cryer PE. Mechanisms of postprandial glucose counterregulation in man. Physiologic roles of glucagon and epinephrine vis-a-vis insulin in the prevention of hypoglycemia late after glucose ingestion. J Clin Invest. 1983;72(1):278-286.
  14. DeFronzo RA. Lilly lecture 1987. The triumvirate: beta-cell, muscle, liver. A collusion responsible for NIDDM. Diabetes. 1988;37(6):667-687.
  15. Unger RH, Cherrington AD. Glucagonocentric restructuring of diabetes: a pathophysiologic and therapeutic makeover. J Clin Invest. 2012;122(1):4-12.
  16. Finan B, Capozzi ME, Campbell JE. Repositioning Glucagon Action in the Physiology and Pharmacology of Diabetes. Diabetes. 2020;69(4):532-541.
  17. Sandoval DA, D'Alessio DA. Physiology of proglucagon peptides: role of glucagon and GLP-1 in health and disease. Physiol Rev. 2015;95(2):513-548.
  18. Towler DA, Havlin CE, Craft S, Cryer P. Mechanism of awareness of hypoglycemia. Perception of neurogenic (predominantly cholinergic) rather than neuroglycopenic symptoms. Diabetes. 1993;42(12):1791-1798.
  19. Whipple AO, Frantz VK. Adenoma of Islet Cells with Hyperinsulinism: A Review. Ann Surg. 1935;101(6):1299-1335.
  20. Mitrakou A, Ryan C, Veneman T, Mokan M, Jenssen T, Kiss I, Durrant J, Cryer P, Gerich J. Hierarchy of glycemic thresholds for counterregulatory hormone secretion, symptoms, and cerebral dysfunction. Am J Physiol. 1991;260(1 Pt 1):E67-74.
  21. Stanley S, Moheet A, Seaquist ER. Central Mechanisms of Glucose Sensing and Counterregulation in Defense of Hypoglycemia. Endocr Rev. 2019;40(3):768-788.
  22. DeRosa MA, Cryer PE. Hypoglycemia and the sympathoadrenal system: neurogenic symptoms are largely the result of sympathetic neural, rather than adrenomedullary, activation. Am J Physiol Endocrinol Metab. 2004;287(1):E32-41.
  23. Lev-Ran A, Anderson RW. The diagnosis of postprandial hypoglycemia. Diabetes. 1981;30(12):996-999.
  24. Charles MA, Hofeldt F, Shackelford A, Waldeck N, Dodson LE, Jr., Bunker D, Coggins JT, Eichner H. Comparison of oral glucose tolerance tests and mixed meals in patients with apparent idiopathic postabsorptive hypoglycemia: absence of hypoglycemia after meals. Diabetes. 1981;30(6):465-470.
  25. Salehi M, Vella A, McLaughlin T, Patti ME. Hypoglycemia After Gastric Bypass Surgery: Current Concepts and Controversies. J Clin Endocrinol Metab. 2018;103(8):2815-2826.
  26. Wang LR, Morein J, McCudden C, Sorisky A. Artifactual hypoglycemia in a patient with sickle cell anemia. CMAJ. 2021;193(43):E1660-E1662.
  27. Cryer PE. Symptoms of hypoglycemia, thresholds for their occurrence, and hypoglycemia unawareness. Endocrinol Metab Clin North Am. 1999;28(3):495-500, v-vi.
  28. Davis SN, Galassetti P, Wasserman DH, Tate D. Effects of antecedent hypoglycemia on subsequent counterregulatory responses to exercise. Diabetes. 2000;49(1):73-81.
  29. Merimee TJ, Tyson JE. Stabilization of plasma glucose during fasting; Normal variations in two separate studies. N Engl J Med. 1974;291(24):1275-1278.
  30. Laurenti MC, Matveyenko A, Vella A. Measurement of Pulsatile Insulin Secretion: Rationale and Methodology. Metabolites. 2021;11(7).
  31. Mittendorfer B, Patterson BW, Smith GI, Yoshino M, Klein S. beta Cell function and plasma insulin clearance in people with obesity and different glycemic status. J Clin Invest. 2022;132(3).
  32. Polonsky KS, Rubenstein AH. C-peptide as a measure of the secretion and hepatic extraction of insulin. Pitfalls and limitations. Diabetes. 1984;33(5):486-494.
  33. Polonsky KS, Pugh W, Jaspan JB, Cohen DM, Karrison T, Tager HS, Rubenstein AH. C-peptide and insulin secretion. Relationship between peripheral concentrations of C-peptide and insulin and their secretion rates in the dog. J Clin Invest. 1984;74(5):1821-1829.
  34. Boden G, Chen X, Desantis RA, Kendrick Z. Effects of insulin on fatty acid reesterification in healthy subjects. Diabetes. 1993;42(11):1588-1593.
  35. Luzi L, Battezzati A, Perseghin G, Bianchi E, Vergani S, Secchi A, La Rocca E, Staudacher C, Spotti D, Ferrari G, et al. Lack of feedback inhibition of insulin secretion in denervated human pancreas. Diabetes. 1992;41(12):1632-1639.
  36. Gin H, Brottier E, Dupuy B, Guillaume D, Ponzo J, Aubertin J. Use of the glucose clamp technique for confirmation of insulinoma autonomous hyperinsulinism. Arch Intern Med. 1987;147(5):985-987.
  37. Nauck MA, Baum F, Seidensticker F, Roder M, Dinesen B, Creutzfeldt W. A hyperinsulinaemic, sequentially eu- and hypoglycaemic clamp test to characterize autonomous insulin secretion in patients with insulinoma. Eur J Clin Invest. 1997;27(2):109-115.
  38. Egan AM, Galior KD, Maus AD, Fatica E, Simha V, Shah P, Singh RJ, Vella A. Pitfalls in Diagnosing Hypoglycemia Due to Exogenous Insulin: Validation and Utility of an Insulin Analog Assay. Mayo Clin Proc. 2022;97(11):1994-2004.
  39. Awad DH, Gokarakonda SB, Ilahi M. Factitious Hypoglycemia. StatPearls. Treasure Island (FL)2023.
  40. Fineberg SE, Kawabata TT, Finco-Kent D, Fountaine RJ, Finch GL, Krasner AS. Immunological responses to exogenous insulin. Endocr Rev. 2007;28(6):625-652.
  41. Oueslati I, Terzi A, Yazidi M, Kamoun E, Chihaoui M. Prevalence and characteristics of factitious hypoglycaemia in non-diabetic patients in a department of endocrinology. Endocrinol Diabetes Metab. 2022;5(6):e375.
  42. Grunberger G, Weiner JL, Silverman R, Taylor S, Gorden P. Factitious hypoglycemia due to surreptitious administration of insulin. Diagnosis, treatment, and long-term follow-up. Ann Intern Med. 1988;108(2):252-257.
  43. Lupsa BC, Chong AY, Cochran EK, Soos MA, Semple RK, Gorden P. Autoimmune forms of hypoglycemia. Medicine (Baltimore). 2009;88(3):141-153.
  44. Cappellani D, Macchia E, Falorni A, Marchetti P. Insulin Autoimmune Syndrome (Hirata Disease): A Comprehensive Review Fifty Years After Its First Description. Diabetes Metab Syndr Obes. 2020;13:963-978.
  45. Jialal I, Basheer H. Syndromes of autoantibodies to the insulin receptor. Int J Biochem Mol Biol. 2022;13(6):87-91.
  46. Zhuo F, Anastasopoulou C. Insulinoma. StatPearls. Treasure Island (FL)2023.
  47. Noone TC, Hosey J, Firat Z, Semelka RC. Imaging and localization of islet-cell tumours of the pancreas on CT and MRI. Best Pract Res Clin Endocrinol Metab. 2005;19(2):195-211.
  48. Giannis D, Moris D, Karachaliou GS, Tsilimigras DI, Karaolanis G, Papalampros A, Felekouras E. Insulinomas: from diagnosis to treatment. A review of the literature. J BUON. 2020;25(3):1302-1314.
  49. Okabayashi T, Shima Y, Sumiyoshi T, Kozuki A, Ito S, Ogawa Y, Kobayashi M, Hanazaki K. Diagnosis and management of insulinoma. World J Gastroenterol. 2013;19(6):829-837.
  50. Placzkowski KA, Vella A, Thompson GB, Grant CS, Reading CC, Charboneau JW, Andrews JC, Lloyd RV, Service FJ. Secular trends in the presentation and management of functioning insulinoma at the Mayo Clinic, 1987-2007. J Clin Endocrinol Metab. 2009;94(4):1069-1073.
  51. Service FJ, Natt N. The prolonged fast. J Clin Endocrinol Metab. 2000;85(11):3973-3974.
  52. Tucker ON, Crotty PL, Conlon KC. The management of insulinoma. Br J Surg. 2006;93(3):264-275.
  53. Goh BK, Ooi LL, Cheow PC, Tan YM, Ong HS, Chung YF, Chow PK, Wong WK, Soo KC. Accurate preoperative localization of insulinomas avoids the need for blind resection and reoperation: analysis of a single institution experience with 17 surgically treated tumors over 19 years. J Gastrointest Surg. 2009;13(6):1071-1077.
  54. Zhao K, Patel N, Kulkarni K, Gross JS, Taslakian B. Essentials of Insulinoma Localization with Selective Arterial Calcium Stimulation and Hepatic Venous Sampling. J Clin Med. 2020;9(10).
  55. Thompson SM, Vella A, Thompson GB, Rumilla KM, Service FJ, Grant CS, Andrews JC. Selective Arterial Calcium Stimulation With Hepatic Venous Sampling Differentiates Insulinoma From Nesidioblastosis. J Clin Endocrinol Metab. 2015;100(11):4189-4197.
  56. Ritzel RA, Isermann B, Schilling T, Knaebel HP, Buchler MW, Nawroth PP. Diagnosis and localization of insulinoma after negative laparotomy by hyperinsulinemic, hypoglycemic clamp and intra-arterial calcium stimulation. Rev Diabet Stud. 2004;1(1):42-46.
  57. Gill GV, Rauf O, MacFarlane IA. Diazoxide treatment for insulinoma: a national UK survey. Postgrad Med J. 1997;73(864):640-641.
  58. Grant CS. Insulinoma. Best Pract Res Clin Gastroenterol. 2005;19(5):783-798.
  59. Mele C, Brunani A, Damascelli B, Ticha V, Castello L, Aimaretti G, Scacchi M, Marzullo P. Non-surgical ablative therapies for inoperable benign insulinoma. J Endocrinol Invest. 2018;41(2):153-162.
  60. Maines E, Urru SAM, Leonardi L, Fancellu E, Campomori A, Piccoli G, Maiorana A, Soffiati M, Franceschi R. Drug-induced hyperinsulinemic hypoglycemia: An update on pathophysiology and treatment. Rev Endocr Metab Disord. 2023.
  61. Paoli A. Ketogenic diet for obesity: friend or foe? Int J Environ Res Public Health. 2014;11(2):2092-2107.
  62. Spoke C, Malaeb S. A Case of Hypoglycemia Associated With the Ketogenic Diet and Alcohol Use. J Endocr Soc. 2020;4(6):bvaa045.
  63. Mebane D, Madison LL. Hypoglycemic Action of Ketones. I. Effects of Ketones on Hepatic Glucose Output and Peripheral Glucose Utilization. J Lab Clin Med. 1964;63:177-192.
  64. Felig P, Cherif A, Minagawa A, Wahren J. Hypoglycemia during prolonged exercise in normal men. N Engl J Med. 1982;306(15):895-900.
  65. Zinker BA, Allison RG, Lacy DB, Wasserman DH. Interaction of exercise, insulin, and hypoglycemia studied using euglycemic and hypoglycemic insulin clamps. Am J Physiol. 1997;272(4 Pt 1):E530-542.
  66. Field JB. Exercise and deficient carbohydrate storage and intake as causes of hypoglycemia. Endocrinol Metab Clin North Am. 1989;18(1):155-161.
  67. Tetzschner R, Norgaard K, Ranjan A. Effects of alcohol on plasma glucose and prevention of alcohol-induced hypoglycemia in type 1 diabetes-A systematic review with GRADE. Diabetes Metab Res Rev. 2018;34(3).
  68. Douillard C, Mention K, Dobbelaere D, Wemeau JL, Saudubray JM, Vantyghem MC. Hypoglycaemia related to inherited metabolic diseases in adults. Orphanet J Rare Dis. 2012;7:26.
  69. Gill RQ, Sterling RK. Acute liver failure. J Clin Gastroenterol. 2001;33(3):191-198.
  70. Mellinkoff SM, Tumulty PA. Hepatic hypoglycemia; its occurrence in congestive heart failure. N Engl J Med. 1952;247(20):745-750.
  71. Miller SI, Wallace RJ, Jr., Musher DM, Septimus EJ, Kohl S, Baughn RE. Hypoglycemia as a manifestation of sepsis. Am J Med. 1980;68(5):649-654.
  72. Maitra SR, Wojnar MM, Lang CH. Alterations in tissue glucose uptake during the hyperglycemic and hypoglycemic phases of sepsis. Shock. 2000;13(5):379-385.
  73. Gosmanov AR, Gosmanova EO, Kovesdy CP. Evaluation and management of diabetic and non-diabetic hypoglycemia in end-stage renal disease. Nephrol Dial Transplant. 2016;31(1):8-15.
  74. Arem R. Hypoglycemia associated with renal failure. Endocrinol Metab Clin North Am. 1989;18(1):103-121.
  75. Husebye ES, Allolio B, Arlt W, Badenhoop K, Bensing S, Betterle C, Falorni A, Gan EH, Hulting AL, Kasperlik-Zaluska A, Kampe O, Lovas K, Meyer G, Pearce SH. Consensus statement on the diagnosis, treatment and follow-up of patients with primary adrenal insufficiency. J Intern Med. 2014;275(2):104-115.
  76. Rushworth RL, Torpy DJ, Falhammar H. Adrenal Crisis. N Engl J Med. 2019;381(9):852-861.
  77. Bornstein SR, Allolio B, Arlt W, Barthel A, Don-Wauchope A, Hammer GD, Husebye ES, Merke DP, Murad MH, Stratakis CA, Torpy DJ. Diagnosis and Treatment of Primary Adrenal Insufficiency: An Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab. 2016;101(2):364-389.
  78. Fukuda I, Hizuka N, Ishikawa Y, Yasumoto K, Murakami Y, Sata A, Morita J, Kurimoto M, Okubo Y, Takano K. Clinical features of insulin-like growth factor-II producing non-islet-cell tumor hypoglycemia. Growth Horm IGF Res. 2006;16(4):211-216.
  79. Tietge UJ, Schofl C, Ocran KW, Wagner S, Boker KH, Brabant G, Zapf J, Manns MP. Hepatoma with severe non-islet cell tumor hypoglycemia. Am J Gastroenterol. 1998;93(6):997-1000.
  80. Schovanek J, Cibickova L, Ctvrtlik F, Tudos Z, Karasek D, Iacobone M, Frysak Z. Hypoglycemia as a Symptom of Neoplastic Disease, with a focus on Insulin-like Growth Factors Producing Tumors. J Cancer. 2019;10(26):6475-6480.
  81. Khowaja A, Johnson-Rabbett B, Bantle J, Moheet A. Hypoglycemia mediated by paraneoplastic production of Insulin like growth factor-2 from a malignant renal solitary fibrous tumor - clinical case and literature review. BMC Endocr Disord. 2014;14:49.
  82. Martens P, Tits J. Approach to the patient with spontaneous hypoglycemia. Eur J Intern Med. 2014;25(5):415-421.
  83. Lages M, Barros R, Moreira P, Guarino MP. Metabolic Effects of an Oral Glucose Tolerance Test Compared to the Mixed Meal Tolerance Tests: A Narrative Review. Nutrients. 2022;14(10).
  84. Holdsworth CD, Turner D, McIntyre N. Pathophysiology of post-gastrectomy hypoglycaemia. Br Med J. 1969;4(5678):257-259.
  85. Capristo E, Panunzi S, De Gaetano A, Spuntarelli V, Bellantone R, Giustacchini P, Birkenfeld AL, Amiel S, Bornstein SR, Raffaelli M, Mingrone G. Incidence of Hypoglycemia After Gastric Bypass vs Sleeve Gastrectomy: A Randomized Trial. J Clin Endocrinol Metab. 2018;103(6):2136-2146.
  86. Goldfine AB, Mun EC, Devine E, Bernier R, Baz-Hecht M, Jones DB, Schneider BE, Holst JJ, Patti ME. Patients with neuroglycopenia after gastric bypass surgery have exaggerated incretin and insulin secretory responses to a mixed meal. J Clin Endocrinol Metab. 2007;92(12):4678-4685.
  87. Salehi M, Gastaldelli A, D'Alessio DA. Blockade of glucagon-like peptide 1 receptor corrects postprandial hypoglycemia after gastric bypass. Gastroenterology. 2014;146(3):669-680 e662.
  88. Lee CJ, Wood GC, Lazo M, Brown TT, Clark JM, Still C, Benotti P. Risk of post-gastric bypass surgery hypoglycemia in nondiabetic individuals: A single center experience. Obesity (Silver Spring). 2016;24(6):1342-1348.
  89. Raverdy V, Baud G, Pigeyre M, Verkindt H, Torres F, Preda C, Thuillier D, Gele P, Vantyghem MC, Caiazzo R, Pattou F. Incidence and Predictive Factors of Postprandial Hyperinsulinemic Hypoglycemia After Roux-en-Y Gastric Bypass: A Five year Longitudinal Study. Ann Surg. 2016;264(5):878-885.
  90. Salehi M, Gastaldelli A, D'Alessio DA. Altered islet function and insulin clearance cause hyperinsulinemia in gastric bypass patients with symptoms of postprandial hypoglycemia. J Clin Endocrinol Metab. 2014;99(6):2008-2017.
  91. Yaqub A, Smith EP, Salehi M. Hyperinsulinemic hypoglycemia after gastric bypass surgery: what's up and what's down? Int J Obes (Lond). 2017.
  92. Honka H, Salehi M. Postprandial hypoglycemia after gastric bypass surgery: from pathogenesis to diagnosis and treatment. Curr Opin Clin Nutr Metab Care. 2019;22(4):295-302.
  93. Lehmann V, Tripyla A, Herzig D, Meier J, Banholzer N, Maritsch M, Zehetner J, Giachino D, Nett P, Feuerriegel S, Wortmann F, Bally L. The impact of postbariatric hypoglycaemia on driving performance: A randomized, single-blind, two-period, crossover study in a driving simulator. Diabetes Obes Metab. 2021;23(9):2189-2193.
  94. Kefurt R, Langer FB, Schindler K, Shakeri-Leidenmuhler S, Ludvik B, Prager G. Hypoglycemia after Roux-En-Y gastric bypass: detection rates of continuous glucose monitoring (CGM) versus mixed meal test. Surg Obes Relat Dis. 2015;11(3):564-569.
  95. Lazar LO, Sapojnikov S, Pines G, Mavor E, Ostrovsky V, Schiller T, Knobler H, Zornitzki T. Symptomatic and Asymptomatic Hypoglycemia Post Three Different Bariatric Procedures: A Common and Severe Complication. Endocr Pract. 2019.
  96. Nguyen NQ, Debreceni TL, Bambrick JE, Bellon M, Wishart J, Standfield S, Rayner CK, Horowitz M. Rapid gastric and intestinal transit is a major determinant of changes in blood glucose, intestinal hormones, glucose absorption and postprandial symptoms after gastric bypass. Obesity (Silver Spring). 2014;22(9):2003-2009.
  97. Patti ME, Goldfine AB. Hypoglycemia after gastric bypass: the dark side of GLP-1. Gastroenterology. 2014;146(3):605-608.
  98. Craig CM, Liu LF, Deacon CF, Holst JJ, McLaughlin TL. Critical role for GLP-1 in symptomatic post-bariatric hypoglycaemia. Diabetologia. 2017;60(3):531-540.
  99. Salehi M, Gastaldelli A, DeFronzo R. Prandial hepatic glucose production during hypoglycemia is altered after gastric bypass surgery and sleeve gastrectomy. Metabolism. 2022:155199.
  100. Salehi M, Woods SC, D'Alessio DA. Gastric bypass alters both glucose-dependent and glucose-independent regulation of islet hormone secretion. Obesity (Silver Spring). 2015;23(10):2046-2052.
  101. Abrahamsson N, Borjesson JL, Sundbom M, Wiklund U, Karlsson FA, Eriksson JW. Gastric Bypass Reduces Symptoms and Hormonal Responses in Hypoglycemia. Diabetes. 2016;65(9):2667-2675.
  102. Salehi M, Gastaldelli A, DeFronzo R. Prandial hepatic glucose production during hypoglycemia is altered after gastric bypass surgery and sleeve gastrectomy. Metabolism. 2022;131:155199.
  103. Honka H, Gastaldelli A, Pezzica S, Peterson R, DeFronzo R, Salehi M. Endogenous glucagon-like peptide 1 diminishes prandial glucose counterregulatory response to hypoglycemia after gastric bypass surgery. medRxiv. 2023.
  104. Axelsen M, Wesslau C, Lonnroth P, Arvidsson Lenner R, Smith U. Bedtime uncooked cornstarch supplement prevents nocturnal hypoglycaemia in intensively treated type 1 diabetes subjects. J Intern Med. 1999;245(3):229-236.
  105. Cummings C, Jiang A, Sheehan A, Ferraz-Bannitz R, Puleio A, Simonson DC, Dreyfuss JM, Patti ME. Continuous glucose monitoring in patients with post-bariatric hypoglycaemia reduces hypoglycaemia and glycaemic variability. Diabetes Obes Metab. 2023;25(8):2191-2202.
  106. Tan M, Lamendola C, Luong R, McLaughlin T, Craig C. Safety, efficacy and pharmacokinetics of repeat subcutaneous dosing of avexitide (exendin 9-39) for treatment of post-bariatric hypoglycaemia. Diabetes Obes Metab. 2020;22(8):1406-1416.
  107. Mulla CM, Zavitsanou S, Laguna Sanz AJ, Pober D, Richardson L, Walcott P, Arora I, Newswanger B, Cummins MJ, Prestrelski SJ, Doyle FJ, Dassau E, Patti ME. A Randomized, Placebo-Controlled Double-Blind Trial of a Closed-Loop Glucagon System for Postbariatric Hypoglycemia. J Clin Endocrinol Metab. 2020;105(4):e1260-1271.
  108. Nielsen CK, Ohrstrom CC, Houji IJK, Helsted MM, Krogh LSL, Johansen NJ, Hartmann B, Holst JJ, Vilsboll T, Knop FK. Dasiglucagon Treatment for Postprandial Hypoglycemia After Gastric Bypass: A Randomized, Double-Blind, Placebo-Controlled Trial. Diabetes Care. 2023;46(12):2208-2217.
  109. Pasireotide s.c. in Patients With Post-Bariatric Hypoglycemia (PASIPHY).
  110. Lawler HM. Inhibition of Intestinal SGLT1 with Mizagliflozin for the Treatment of Post-bariatric Hypoglycemia J Endocr Soc. 2023;7:A446.
  111. Kloppel G, Anlauf M, Raffel A, Perren A, Knoefel WT. Adult diffuse nesidioblastosis: genetically or environmentally induced? Hum Pathol. 2008;39(1):3-8.
  112. Dravecka I, Lazurova I. Nesidioblastosis in adults. Neoplasma. 2014;61(3):252-256.
  113. Service FJ, Natt N, Thompson GB, Grant CS, van Heerden JA, Andrews JC, Lorenz E, Terzic A, Lloyd RV. Noninsulinoma pancreatogenous hypoglycemia: a novel syndrome of hyperinsulinemic hypoglycemia in adults independent of mutations in Kir6.2 and SUR1 genes. J Clin Endocrinol Metab. 1999;84(5):1582-1589.
  114. Thompson GB, Service FJ, Andrews JC, Lloyd RV, Natt N, van Heerden JA, Grant CS. Noninsulinoma pancreatogenous hypoglycemia syndrome: an update in 10 surgically treated patients. Surgery. 2000;128(6):937-944;discussion 944-935.
  115. Won JG, Tseng HS, Yang AH, Tang KT, Jap TS, Lee CH, Lin HD, Burcus N, Pittenger G, Vinik A. Clinical features and morphological characterization of 10 patients with noninsulinoma pancreatogenous hypoglycaemia syndrome (NIPHS). Clin Endocrinol (Oxf). 2006;65(5):566-578.
  116. Witteles RM, Straus IF, Sugg SL, Koka MR, Costa EA, Kaplan EL. Adult-onset nesidioblastosis causing hypoglycemia: an important clinical entity and continuing treatment dilemma. Arch Surg. 2001;136(6):656-663.
  117. Then C, Nam-Apostolopoulos YC, Seissler J, Lechner A. Refractory idiopathic non-insulinoma pancreatogenous hypoglycemia in an adult: case report and review of the literature. JOP. 2013;14(3):264-268.
  118. Vanderveen KA, Grant CS, Thompson GB, Farley DR, Richards ML, Vella A, Vollrath B, Service FJ. Outcomes and quality of life after partial pancreatectomy for noninsulinoma pancreatogenous hypoglycemia from diffuse islet cell disease. Surgery. 2010;148(6):1237-1245; discussion 1245-1236.

 

Triglyceride Lowering Drugs

ABSTRACT

 

The two major goals of the treatment of hypertriglyceridemia are the prevention of cardiovascular disease and pancreatitis. Here we discuss the drugs used for the treatment of hypertriglyceridemia: (niacin, fibrates, omega-3-fatty acids, volanesorsen (available in Europe) and lipoprotein lipase gene therapy (alipogene tiparvovec- no longer available). Niacin decreases total cholesterol, TGs (20-50% decrease), LDL-C, and Lp(a). Additionally, niacin decreases small dense LDL resulting in a shift to large, buoyant LDL particles. Moreover, niacin increases HDL-C. Skin flushing, insulin resistance, and other side effects have limited the use of niacin. The enthusiasm for niacin has greatly decreased with the failure of AIM-HIGH and HPS-2 Thrive to decrease cardiovascular events when niacin was added to statin therapy. The omega-3-fatty acids eicosapentaenoic acid (C20:5n-3) (EPA) and docosahexaenoic acid (C22:6n-3) (DHA) lower TGs by 10-50% but do not affect total cholesterol, HDL-C, or Lp(a). LDL-C may increase with EPA + DHA when the TG levels are markedly elevated (>500mg/dL). EPA alone does not increase LDL-C. Omega-3-fatty acids have few side effects, drug interactions, or contraindications. Numerous studies of low dose omega-3-fatty acids on cardiovascular outcomes have failed to demonstrate a benefit. However, in the JELIS trial and REDUCE-IT trial high doses of EPA alone reduced cardiovascular events while in the STRENGTH trial high dose EPA+DHA did not reduce cardiovascular events. Fibrates reduce TG levels by 25-50% and increase HDL-C by 5-20%. The effect on LDL-C is variable. If the TG levels are very high (>500mg/dL), fibrate therapy may result in an increase in LDL-C, whereas if TGs are not markedly elevated fibrates decrease LDL-C by 10-30%. Fibrates also reduce apolipoprotein B, LDL particle number, and non-HDL-C and there may be a shift from small dense LDL towards large LDL particles. Fibrates do not have any major effects on Lp(a). Monotherapy with fibrates appears to reduce cardiovascular events in patients with high TG and low HDL-C levels. Whether the addition of fibrates to statin therapy will reduce cardiovascular disease is uncertain. In patients with diabetes fibrates appear to slow the progression of microvascular disease. Volanesorsen is an antisense oligonucleotide that inhibits the production of apolipoprotein C-III. In patients with the familial chylomicronemia syndrome (FCS) volanesorsen decreases TG by 77% (mean decrease of 1712 mg/dL) with 77% of the patients having TG levels less than 750 mg/dL. In addition, volanesorsen treatment resulted in decreases in non–HDL-C by 46%, and VLDL-C by 58% and increases in HDL-C by 46%, LDL-C by 136%, (LDL-C increased from 28 to 61 mg/dL), and total apolipoprotein B by 20%. Studies have suggested that volanesorsen may reduce episodes of pancreatitis. Patients with FCS have also reported that volanesorsen improved symptoms and reduced interference of FCS with work/school responsibilities. Of concern has been decreases in platelet levels with 47% of patients treated with volanesorsen developing platelet counts below100 x 109/L. Thus, a number of drugs are available for the treatment of hypertriglyceridemia and may be employed when lifestyle changes are not sufficient. 

 

INTRODUCTION

 

The two primary goals of the treatment of hypertriglyceridemia are the prevention of cardiovascular disease and the prevention of pancreatitis. The evaluation and guidelines for the management of hypertriglyceridemia are discussed in detail in the Endotext chapter “Risk of Fasting and Non-Fasting Hypertriglyceridemia in Coronary Vascular Disease and Pancreatitis” (1) and the approach to evaluating a patient with hypertriglyceridemia is discussed in the Endotext chapter “Approach to the Patient with Dyslipidemia” (2). The treatment of hypertriglyceridemia by diet and weight loss are discussed in detail in the Endotext chapter “The Effect of Diet on Cardiovascular Disease and Lipid and Lipoprotein Levels” and “Obesity and Dyslipidemia” (3,4). Lifestyle changes are recommended as the first line for therapy of hypertriglyceridemia, but drug therapy is often required. In this chapter we will discuss the drugs used for the treatment of elevated plasma TG levels. Statins, ezetimibe, PCSK9 inhibitors, bempedoic acid, lomitapide, mipomersen, and evinacumab, which are primarily used to lower LDL-C, are discussed in the chapter “Cholesterol Lowering Drugs” (5).  

 

NIACIN

 

Introduction

 

Niacin was the first drug approved to treat dyslipidemia. In 1955, Altschul et al showed that pharmacologic doses of niacin decreased plasma cholesterol levels (6). Several forms of niacin are available for clinical use. Immediate release niacin has a short duration of action and is typically given two or three times per day with meals, whereas sustained release niacin and extended-release niacin are once a day drugs usually given at bedtime. The extended release form of niacin exhibits release rates that are intermediate between immediate release niacin and sustained release niacin (7). While the effects of the various forms of niacin on plasma lipid levels are similar, the side effect profiles are different. Because of an increased risk of serious liver toxicity with sustained release niacin this preparation is no longer widely used to treat dyslipidemia. Over-the- counter “No flush” niacin is also available but is generally ineffective as a lipid-modifying agent because most of these preparations do not contain active nicotinic acid.    

 

Effect of Niacin on Lipid and Lipoprotein Levels

 

Table 1. Effect of Niacin on Lipid and Lipoproteins

Decreases Total Cholesterol

Decreases LDL-C

Decreases TGs

Decreases Non-HDL-C

Decreases Lp(a)

Increases HDL-C

Decreases Apolipoprotein B

Shifts Small Dense LDL to Large Buoyant LDL

 

Niacin decreases all the pro-atherogenic lipid and lipoprotein particles including total cholesterol, TG, LDL-C, and Lp(a) levels (Table 1) (8,9). Additionally, niacin has been shown to decrease small dense LDL resulting in a shift to large, buoyant LDL particles (10). Moreover, niacin increases HDL-C levels (8,9).

 

In a meta-analysis of 30 trials with 4,749 subjects treatment with immediate release, sustained release, or extended release niacin decreased total cholesterol by 10%, decreased TGs by 20%, decreased LDL-C by 14%, and increased HDL-C by 16% (11). All three niacin preparations were effective in decreasing total cholesterol, TG, and LDL-C levels and increasing HDL-C levels (11). At a dose of 1.5 grams per day, immediate release niacin and extended release niacin produced similar decreases in total cholesterol, TGs, and LDL-C and a similar increase in HDL-C (12). A meta-analysis of 14 studies with 9,013 subjects reported a 23% decrease in Lp(a) with extended release niacin treatment (13).

 

A small meta-analysis of 5 trials in 432 subjects compared the response to extended release niacin in men and women (14). The effect of niacin on LDL-C was greater in women than men at all niacin doses (1,000mg 6.8% decrease in women vs 0.2% in men, p = 0.006; 1,500mg 11.3% decrease vs 5.6% decrease, p = 0.013; 2,000 mg 14.8% decrease vs 6.9% decrease, p = 0.010; 3,000mg 28.7% decrease vs 17.7% decrease, p = 0.006). The effect of niacin on plasma TG levels also tended to be greater in women but the difference only reached statistical significance at the 1,500mg dose (28.6% vs 20.4%, p = 0.040). The mechanism for the more robust decrease in LDL-C and TGs in women is unknown but might be due to a smaller body mass in women leading to increased circulating niacin levels and hence a greater response. However, the effect of niacin on HDL-C and Lp(a) levels were similar in males and females. Not unexpectedly the effect of niacin is dose dependent with higher doses having a greater effect on plasma lipid and lipoprotein levels (Table 2) (14).

 

Table 2. Effect of Niacin Dose on Lipid and Lipoprotein Response in Women (percent change)

Niacin Dose

LDL-C

TG

HDL-C

Lp(a)

500mg

-5.2

-9.5

7.7

-2.6

1000mg

-6.8

-14.5

17.6

-11.5

1500mg

-11.3

-28.6

21.1

-4.0

2000mg

-14.8

-37.3

25.2

-24.7

2500mg

-28.7

-45.6

34.5

-28.6

3000mg

-28.7

-51.0

28.7

-29.9

 

Numerous studies have examined the effect of the addition of niacin to statin therapy. Combination therapy typically results in further reductions in atherogenic lipoprotein particles and an increase in HDL-C levels. An example of such a study is shown in Table 3 (15).

 

Table 3. Effect of the Addition of Niacin to Statin Therapy on Lipid and Lipoprotein Levels

 

 Change in Lipids with Addition of Extended-Release Niacin 2000mg/day to Simvastatin 20mg/day

LDL-C

7.1% Decrease

HDL-C

18.2% Increase

TG

22.7% Decrease

Non-HDL-C

15.1% Decrease

Lp(a)

17.4% Decrease

 

While a literature search did not find any studies comparing the combination of ezetimibe + niacin vs. monotherapy there is a large trial that has examined the effect of adding 2 grams niacin to ezetimibe/simvastatin 10/20 (16). In this study the addition of niacin improved the lipid profile with a marked decrease in TGs and an increase in HDL-C levels (table 4).

 

Table 4. Effect of the Addition of Niacin to Ezetimibe/Statin Therapy on Lipid and Lipoprotein Levels

 

 Change in Lipids with Addition of Niacin 2000mg/day to Ezetimibe/Simvastatin 10/20mg/day

LDL-C

4.8% Decrease

HDL-C

21.5% Increase

TG

17.6% Decrease

Non-HDL-C

7.3% Decrease

 

In patients with marked hypertriglyceridemia combining niacin with other drugs that also lower plasma TGs can be considered. Sixty patients with the metabolic syndrome were randomized to 16 weeks of treatment with placebo, omega-3-fatty acids (Lovaza 4 g/day), extended release niacin (2 g/day), or both drugs in combination (17). In the niacin group TGs were decreased by 30%, in the omega-3-fatty acids group by 22%, and in the combination group by 42% compared to the placebo group. Of note the beneficial effects of niacin on decreasing LDL and non-HDL-C levels were blunted by omega-3-fatty acids, which are known to raise LDL-C levels in patients with marked hypertriglyceridemia (see below). These results show that the combination of niacin and fish oil will lower TG levels more than either drug individually but at the expense of diminishing the effect of niacin on LDL and non-HDL-C levels.   

 

Surprisingly there are few large randomized trials examining the effect of combination therapy with niacin + fibrate vs. monotherapy. One very small trial reported that while both niacin monotherapy and bezafibrate monotherapy were effective in lowering serum TGs there was no statistically significant added benefit of combination therapy in reducing serum TG levels (18). However, a larger trial in HIV+ patients reported that the combination of niacin and fenofibrate was better at lowering TGs and non-HDL-C and increasing HDL-C levels than monotherapy with either niacin or fenofibrate (19). It would be informative if additional trials of combination therapy were carried out in patients with marked hypertriglyceridemia that can often be difficult to control with lifestyle changes and monotherapy.

 

Mechanisms Accounting for the Niacin Induced Lipid Effects

 

TRIGLYCERIDES

 

Early studies demonstrated that niacin inhibited the release of free fatty acids from cultured adipocytes and decreased circulating free fatty acid levels (20-22). The ability of niacin to inhibit adipose tissue lipolysis is mediated by the activation of GPR109A (hydroxycarboxylic acid 2 receptor), a G protein-coupled receptor that is highly expressed in adipose tissue (22-24). It was initially thought that the decrease in plasma TGs induced by niacin therapy was due to niacin inhibiting lipolysis in adipose tissue resulting in a decrease in the transport of fatty acids to the liver leading to the decreased availability of fatty acids for hepatic TG synthesis. However, studies have shown that while niacin acutely decreases plasma free fatty acid levels this inhibition is not sustained (25). Additionally, studies in mice lacking GPR109A have shown that niacin does not inhibit lipolysis but still decreases plasma TG and LDL-C levels (26). Moreover, studies in humans using GPR109A agonists lowered plasma free fatty acid levels but did not cause the expected effects on plasma TGs and LDL-C (26). Thus, the effects of niacin on adipose tissue lipolysis are no longer thought to mediate the niacin induced decrease in plasma TG levels.

 

Niacin has been shown to inhibit diglycerol acyltransferase 2 (DGAT2) activity in the liver (22,27). DGAT2 is the key enzyme that catalyzes the final step in TG synthesis. Inhibition of DGAT2 will reduce hepatic TG synthesis and the availability of TG for VLDL assembly and secretion (22). A decrease in TG will result in an increase in apolipoprotein B degradation in the liver. Kinetic studies in humans have shown that treatment with niacin decreases VLDL TG production (28,29).

 

In addition, in animal models, niacin reduces the hepatic expression of apolipoprotein C-III, which could result in the accelerated clearance of TG rich lipoproteins (30). Whether this plays a significant role in mediating the decrease in plasma TG levels induced by niacin therapy remains to be determined.

 

LOW DENSITY LIPOPROTEIN

 

The decrease in plasma LDL-C with niacin therapy is thought to be secondary to a reduction in VLDL and LDL formation and secretion by the liver (22).

 

HIGH DENSITY LIPOPROTEIN

 

There are multiple potential mechanisms by which niacin may increase HDL-C levels. Some of these changes may be anti-atherogenic while others may be pro-atherogenic. One hypothesis for the increase in HDL induced by niacin therapy is a decrease in the surface expression of hepatocyte beta chain ATP synthase, a receptor that has been proposed to be involved in the uptake of HDL particles by the liver (31). Studies have further shown that niacin inhibits HDL protein degradation by cultured hepatocytes but does not inhibit the selective uptake of cholesterol esters carried in HDL (22,32).

 

Some kinetic studies have shown that niacin decreases HDL and apolipoprotein A1 fractional catabolic rate (33,34). In contrast, other kinetic studies have shown that niacin increase apolipoprotein AI production (35).

 

In addition, in monocytes, niacin also increased the expression of ABCA1 and CD36 resulting in an increase in cholesterol efflux to HDL, which would increase HDL-C levels and likely have anti-atherogenic effects (36). Similarly, in vitro studies suggest that niacin may increase the transport of cholesterol and phospholipids via ABCA1 from the liver to lipid poor apolipoprotein A1 particles thereby decreasing the clearance of apolipoprotein A1, which might not be anti-atherogenic (22,37).

 

Finally, decreasing plasma TG levels may result in a reduction in CETP mediated exchange of TGs on VLDL for cholesterol on HDL leading to an increase in HDL-C levels. Additionally, studies have shown that niacin decreases the expression of CETP (38).

   

LIPOPROTEIN(a)

 

Niacin decreases the synthetic rate of Lp(a) but does not increase Lp(a) catabolism (39,40). In cell culture and animal studies niacin has been shown to decrease the expression of apo (a) (41).

 

Pharmacokinetics

 

Oral niacin is well absorbed with immediate release niacin resulting in a rapid increase in plasma levels while extended release and sustained release niacin result in a delayed peak in plasma levels. Niacin undergoes metabolism in the liver by two primary pathways; conjugation or amidation (7,42). The conjugative pathway is low affinity and high capacity that metabolizes niacin to nicotinuric acid while the amidation pathway is high affinity and low capacity that converts niacin into several oxidative-reductive intermediates, which can induce hepatic toxicity (7,42) (Figure 1). The clinical importance is that immediate release niacin results in high levels of niacin and therefore is primarily metabolized by the conjugative pathway (low affinity, high capacity), which does not result in toxic intermediates that can cause liver damage. In contrast, sustained release niacin results in lower levels of niacin for a longer period and therefore metabolism via the amidation pathway (high affinity, low capacity) is dominant leading to an increase in the formation of toxic intermediates that can induce hepatic injury (7,42). Extended-release niacin would be metabolized midway between immediate release and sustained release niacin (42).

 

Figure 1. Pathways of Niacin Metabolism.

 

Effect of Niacin on Cardiovascular Outcomes

 

MONOTHERAPY

 

The Coronary Drug Project, conducted between 1966 and 1975, was the first large randomized, double-blind clinical trial to show that lowering lipids reduced cardiovascular disease (43). This trial determined the effect of clofibrate (1.8g/day), dextrothyroxine (6mg/day), two doses of oral estrogen (2.5 or 5mg per day), or immediate release niacin (3 grams/day) vs. placebo in 8,341 men 30 to 64 years of age with an electrocardiogram documented myocardial infarction. The mean baseline total cholesterol level was 251mg/dL and the TG level was 183mg/dL. The two estrogen regimens and dextrothyroxine treatment groups were discontinued early because of increased adverse effects. Clofibrate treatment did not demonstrate clinical benefit. In the niacin treated patients there was an average 10% decrease in serum cholesterol and 26% decrease in serum TGs despite modest compliance with the study medication. Moreover, niacin treatment (n=1,119) decreased recurrent myocardial infarctions by 26%, stroke by 24%, and revascularization by 67% compared to placebo (n=2,789) but did not decrease total mortality, which was the primary endpoint. Long term follow-up (6.2 years during the study and 8.8 years post study after niacin was discontinued in most participants) demonstrated an 11% decrease in mortality in the niacin group vs. the placebo group (52.0 versus 58.2%; p = 0.0004) (44). The majority of this difference in mortality was accounted for by a decrease in coronary heart disease mortality (36.5% vs. 41.3%; p=0.005). Further analysis revealed that niacin reduced the risk of 6-year recurrent myocardial infarction and coronary heart disease death and 15-year total mortality similarly in patients at all levels of baseline fasting plasma glucose, including those with glucose levels ≥126mg/dL (i.e. patients with diabetes) (45). Additionally, the beneficial effect of niacin on cardiovascular events and total mortality was not diminished, even among those with one hour plasma glucose levels > 220mg/dL (45). Moreover, the beneficial effects of niacin on recurrent myocardial infarction and total mortality were similar in patients with or without the metabolic syndrome at baseline (46). These results demonstrate that immediate release niacin monotherapy decreases recurrent atherosclerotic cardiovascular events in a broad spectrum of patients with pre-existing cardiovascular disease (secondary prevention).

 

COMBINATION WITH FIBRATES

 

In the Stockholm Ischemic Heart Disease Secondary Prevention Study survivors of a myocardial infarction below 70 years of age were randomized to a control group (n = 276) (no placebo) and a group treated with clofibrate (2 grams) and immediate release nicotinic acid (up to 3 grams) (n = 279) (47). Serum cholesterol and TG was lowered by 13% and 19%, respectively, in the treatment group compared to the control group. Recurrent myocardial infarction was reduced by 50% within one year (48). Total mortality was decreased by 26% in the group treated with clofibrate + niacin (p< 0.05) while ischemic heart disease mortality was decreased by 36% (p< 0.01). Notably, the benefit of clofibrate + niacin was only observed in patients with a baseline TG level > 143mg/dL. In the age of statins, the clinical implications of this early study are unclear. 

 

COMBINATION WITH STATINS

 

The AIM-HIGH trial was designed to determine if the addition of Niaspan, an extended-release form of niacin, to aggressive statin therapy would result in a further reduction in cardiovascular events in patients with pre-existing cardiovascular disease (49). In this trial 3,314 patients were randomized to extended-release Niaspan (1500-2000mg/day) vs. placebo that contained 100-150mg of immediate release niacin. On trial, LDL-C levels were in the 60-70mg/dL range in both groups. As expected, HDL-C levels were increased in the Niaspan treated group (approximately 44mg/dL vs. 38mg/dL), while TGs were decreased (approximately 121mg/dL vs. 155mg/dL). However, there were no differences in the primary endpoint between the control and Niaspan treated groups (Primary endpoint consisted of the first event of death from coronary heart disease, nonfatal myocardial infarction, ischemic stroke, hospitalization for an acute coronary syndrome, or symptom-driven coronary or cerebral revascularization). There were also no differences in secondary endpoints except for a possible increase in strokes in the Niaspan treated group. The addition of Niaspan to statin therapy did not result in a significant increase in either muscle or liver toxicity. Thus, this study does not provide support for the addition of niacin to statins. However, most of the patients included in this study did not have a lipid profile that one would typically consider treating with niacin therapy. In the subset of patients with TG > 198mg/dL and HDL-C < 33mg/dL Niaspan treatment showed a trend towards benefit (hazard ratio 0.74; p=0.073), suggesting that if the appropriate patient population was studied the results may have been different (50).

 

HPS 2 Thrive also studied the effect of niacin added to statin therapy (51). This trial utilized extended-release niacin (2000mg/day) combined with laropiprant, a prostaglandin D2 receptor antagonist, which reduces the flushing side effect of niacin treatment. HPS 2 Thrive was a very large trial with over 25,000 patients randomized to either niacin therapy or placebo. As in the AIM HIGH study, the baseline LDL-C levels were low at 63mg/dL, the HDL-C levels were 44mg/dL, and the TGs were 125mg/dL at baseline. As expected, niacin therapy resulted in a modest reduction in LDL-C (10mg/dL), a modest increase in HDL-C (6mg/dL), and a marked reduction in TGs (33mg/dL) compared to placebo. However, despite these lipid changes there were no significant differences in major cardiovascular events between the niacin and control group (risk ratio 0.96 CI 0.90- 1.03). It is unknown whether laropiprant, the prostaglandin D2 receptor antagonist, might have effects that worsen atherosclerosis and increase event rates. Mice deficient in the prostaglandin D2 receptor have been noted to have an increase in atherogenesis in response to angiotensin II (52). Similar to the AIM-HIGH study, the group of patients included in the HPS 2 Thrive trial may not have been the ideal patient population to study for the beneficial effects of niacin treatment added to statin therapy. Ideally, patients with high TGs and high non-HDL-C levels coupled with low HDL-C levels should be studied.

 

Thus, these two studies have failed to demonstrate that adding niacin to statin therapy results in a decrease in cardiovascular events. It should be recognized that both the AIM-HIGH study and the HPS-2 Thrive study had limitations. First, the patient populations that were included in these studies were not ideal as the TG and non-HDL-C levels were not elevated in a range that one would usually consider adding niacin therapy. Second, in both trials a significant percentage of patients stopped niacin therapy (AIM-HIGH 25.4% discontinued niacin; HPS-2 Thrive 25.4% discontinued niacin). Third, the duration of these studies was relatively short and it is possible that the beneficial effects of niacin take longer to occur (AIM-HIGH 3 years; HPS-2 Thrive 3.9 years). Fourth, in the HPS-2 Thrive it is possible, as noted earlier, that laropiprant had adverse effects that increased the risk of cardiovascular events. Fifth, in the AIM-HIGH study the placebo contained a low dose of niacin, which may have resulted in beneficial effects. Finally, both of these trials used extended-release niacin, whereas the Coronary Drug Project and the Stockholm Ischemic Heart Disease Secondary Prevention Study used immediate release niacin. It is possible that these different formulations of niacin have different effects on cardiovascular events. Additional studies are required to definitively determine the effect of niacin added to a statin therapy on cardiovascular events.

 

Effect of Niacin on Atherosclerosis

 

Many of the initial niacin therapy imaging studies combined niacin with other drugs and compared these combinations vs. placebo. These studies showed that niacin in combination with other drugs reduced the progression and/or increased the regression of atherosclerosis. However, because of the use of other drugs it is impossible to determine if niacin therapy per se was beneficial (Table 5).

 

Table 5. Niacin Angiography Imaging Studies

Combination Studies

Drugs

Cholesterol Lowering Atherosclerosis Study (CLAS) (53)

Niacin + colestipol vs. placebo

Familial Atherosclerosis Treatment Study (FATS) (54)

Niacin + colestipol or lovastatin + colestipol vs. placebo

UCSF-SCORE (55)

Niacin + colestipol +/- lovastatin vs. placebo +/- low dose colestipol

HDL Atherosclerosis Study (HATS) (56)

Niacin + simvastatin vs. placebo

Armed Forces Regression Study (57)

Niacin + gemfibrozil + cholestyramine vs. placebo

Harvard Atherosclerosis Reversibility Project (HARP)  (58)

Niacin + pravastatin + cholestyramine + gemfibrozil as needed vs. placebo

 

However, there are studies that compared niacin to placebo or other drugs added to standard statin therapy that do provide useful insights (Table 6).

 

Table 6. Effect of Niacin Added to Statin Therapy on Atherosclerosis

ARBITER 2/3

(59,60)

ER niacin vs. placebo

Decrease in CIMT vs. placebo

ARBITER 6 (61)

ER niacin vs. ezetimibe

Decrease in CIMT vs. ezetimibe

Thoenes (62)

ER niacin vs. placebo

Decrease in CIMT vs. placebo

Lee (63)

Modified release niacin vs. placebo

Decrease in carotid wall area on MRI vs. placebo

 

The ARBITER 2 Trial was a double-blind randomized study of extended-release niacin (1000mg) vs. placebo added to background statin therapy in 167 patients with coronary heart disease and low HDL-C levels (<45mg/dL) (60). At the initiation of the study mean LDL-C levels were < 100mg/dL. The primary end point was the change in common carotid intima-media thickness (CIMT). As expected, plasma TGs decreased and HDL-C levels increased with niacin therapy. LDL-C levels were unchanged. After 12 months, mean CIMT increased significantly in the placebo group (P<0.001) and was unchanged in the niacin group (P=0.23). The overall difference in CIMT progression between the niacin and placebo groups was almost statistically significant (P=0.08). Cardiovascular events occurred in 3 patients treated with niacin (3.8%) and 7 patients treated with placebo (9.6%; P=0.20). ARBITER 3 was a 12-month extension and in the 57 patients that continued on niacin therapy there was an additional regression of CIMT (p = 0.001 vs. placebo) (59).

 

In ARBITER 6, patients with coronary heart disease or a coronary heart disease risk equivalent on long-term statin therapy with LDL-C level < 100mg/dL and an HDL-C level < 50mg/dL for men or 55mg/dL for women were randomly assigned to receive either extended-release niacin (target dose, 2000mg per day) or ezetimibe (10mg per day) (61). The primary end point was the change from baseline in the mean CIMT. LDL-C levels decreased in the ezetimibe group by −18mg/dL (~ 20%) and by −10.0mgdl (~ 12%) in the niacin group (P=0.01) while HDL-C levels were slightly decreased in the ezetimibe group −2.8mg/dL and increased by 7.5mg/dL (~18%) in the niacin group (P<0.001). TG levels were not markedly altered in the ezetimibe group but decreased by ~ 15-20% in the niacin group.  Notably niacin therapy resulted in a significant reduction of both mean (P = 0.001) and maximal CIMT (P < 0.001) while ezetimibe therapy significantly increased CIMT (P < 0.001). The incidence of major cardiovascular events was lower in the niacin group than in the ezetimibe group (1% vs. 5%, P = 0.04).

 

In a trial by Thoenes and colleagues fifty patients with the metabolic syndrome not on statin therapy were randomized to either extended-release niacin (1000mg/day) or placebo (62). Treatment with niacin decreased LDL-C by 17% and TGs by 23% and increased HDL-C levels by 24% without significant changes in the placebo group. After 52 weeks of treatment, there was an increase in CIMT of +0.009 +/- 0.003 mm in the placebo group and a decrease in CIMT of -0.005 +/- 0.002 mm in the niacin group (p = 0.021 between groups).

 

Finally, Lee and colleagues performed a double-blind, randomized study of 2 g daily modified-release niacin or placebo added to statin therapy in 71 patients with low HDL-C (<40mg/dL) and either: 1) type 2 diabetes with coronary heart disease; or 2) carotid/peripheral atherosclerosis (63). The primary end point was the change in carotid artery wall area, quantified by magnetic resonance imaging, after 1 year. Treatment with niacin increased HDL-C by 23% and decreased LDL-C by 19% and TGs by 11%. At 12 months, niacin significantly reduced carotid wall area compared with placebo (Mean change in carotid wall area was -1.1 +/- 2.6 mm2 for niacin vs +1.2 +/- 3.0 mm2 for placebo).

 

While these imaging studies provide data suggesting that niacin therapy when added to statin therapy may reduce atherosclerotic cardiovascular disease, one must recognize that the studies described above were relatively small studies and that decreases or the lack of progression in CIMT or carotid wall area are surrogate markers, which may not necessarily indicate that cardiovascular events will be decreased.  

 

Side Effects

 

Treatment with niacin frequently results in side effects and these side effects are a major limitation of niacin therapy.

 

SKIN FLUSHING

 

This is a very common side effect and is characterized by redness and warmth due to vasodilation of the blood vessels in the skin (8,64). It is often most apparent in the head and neck region. Itching can occur and a tingling and burning sensation may also be noted. Niacin induced flushing is usually not accompanied by diaphoresis. The cutaneous flushing usually lasts for approximately one hour and in some patients is extremely annoying. In a review of 30 studies, it was noted that flushing occurred in 85% of participants treated with immediate release niacin, 66% of participants treated with extended release niacin, and 26% of participants treated with slow release niacin (11).  The occurrence of flushing is related to a rapid increase in plasma nicotinic acid levels, which differs depending upon the niacin preparation. Flushing was the primary reason that subjects discontinued niacin therapy during studies and with either immediate release or extended release niacin approximately 20% of study participants discontinue niacin, which is twice the rate of discontinuation observed in the placebo groups (11). Continuous administration of niacin for approximately one- week results in tachyphylaxis and the flushing decreases. Unfortunately, if a patient skips taking niacin for a few days this tachyphylaxis is lost and the flushing returns.

 

The mechanism for the niacin induced skin flushing has been partially elucidated (8,64). Niacin activates GPR109A in dermal Langerhan cells (macrophages in the skin), which leads to the increased production of prostaglandin D2.  Additionally, niacin activates GPR 109A in keratinocytes, which leads to the production of prostaglandin E2.  The prostaglandins then interact with prostaglandin receptors on blood vessels resulting in vasodilation and the flushing phenomena. Aspirin and nonsteroidal anti-inflammatory drugs (NSAIDS) taken prior to niacin administration can decrease flushing by inhibiting the synthesis of prostaglandins (8,65). Laropiprant decreases flushing by blocking the D prostanoid receptor (8). Since flushing is related to rapid increases in plasma nicotinamide levels taking immediate release niacin with food slows absorption and thereby reduces flushing. Extended-release niacin is typically taken at bedtime so that the flushing will occur when the patient is asleep. Conditions that predispose to cutaneous vasodilatation such as alcohol intake, hot liquids, spicy foods, or hot showers should be avoided. One should increase the dose of niacin slowly to reduce the severity of flushing reactions and allow tolerance to develop.

 

HEPATIC TOXICITY

 

Sustained release niacin has a much greater propensity to induce hepatic toxicity than other niacin preparations and therefore is no longer widely used (7,42,66). The explanation for this difference is due to the increased metabolism of sustained release niacin by the amidation pathway described in the pharmacokinetics section, which results in toxic compounds that injure the liver (7,42). Patients who have developed signs of liver toxicity on sustained release niacin can often be treated with immediate release niacin without developing liver problems (67). Extended-release niacin can induce liver dysfunction but the rate is much lower than sustained release niacin. Because of the potential for liver disease, serum transaminase levels (SGOT and SGPT) should be monitored before treatment begins, every 6 to 12 weeks for the first year, and periodically thereafter (e.g., at approximately 6-month intervals).

 

It should be noted that there is some evidence that niacin may be beneficial for non-alcoholic fatty liver disease (NAFLD) but further studies are required (68).

 

MUSCLE SYMPTOMS

 

Myalgias and myopathy have not been a significant adverse effect with niacin monotherapy (11). In combination with statins, an increased risk of muscle symptoms has been observed in some studies. In the HPS-2 Thrive study the combination of simvastatin and extended-release niacin increased the risk of myopathy four-fold (1.2% of patients on combined therapy) (51). Of note, this increase occurred predominantly in Chinese participants. In contrast, in the AIM-HIGH trial muscle related symptoms were not increased with the simvastatin + niacin combination (49,69).

 

HYPERGLYCEMIA

 

It has been recognized for many years that niacin induces insulin resistance (70). The mechanisms by which niacin induces insulin resistance are unknown but possible mechanisms include a rebound increase in free fatty acids with niacin therapy or the accumulation of diacylglycerol (29,71). A recent analysis of the AIM-HIGH trial demonstrated that in subjects with normal glucose metabolism, subjects with impaired fasting glucose, and subjects with diabetes, treatment with extended release niacin resulted in only small increases in fasting glucose levels but increased serum insulin levels due to an increase in insulin resistance (72). Additionally, there was an increased risk of progressing from normal to impaired fasting glucose in subjects treated with niacin in the AIM-HIGH trial (niacin 58.6% vs placebo 41.5%; P < .001) (72).

 

A meta-analysis examined the effect of niacin therapy on the development of new onset diabetes (73). In 11 trials with 26,340 non-diabetic participants, niacin therapy was associated with a 34% increased risk of developing diabetes (RR of 1.34; 95% CIs 1.21 to 1.49). This increased risk results in one additional case of diabetes per 43 initially non-diabetic individuals who are treated with niacin for 5 years (0.47% ten-year risk or 4.7 per 1000 patient years). Results were similar in patients who were receiving niacin therapy in combination with statin therapy.

 

Studies have shown that niacin is usually well tolerated in diabetic subjects who are in good glycemic control (74,75). In patients with poor glycemic control, niacin is more likely to adversely impact glucose levels. A meta-analysis of 7 studies with 838 patients with diabetes found that niacin therapy did not result in a significant increase in fasting glucose levels in short term studies but in long term studies there was a very small increase in fasting glucose levels (1.5mg/dL) that was not clinically significant (76). An important caveat is that in most of these trials adjustments in diabetes therapy was permitted, which could blunt worsening of glycemic control. In contrast to these findings, the HPS-2 Thrive Trial reported that in the 8,299 participants who had diabetes at the time of randomization, treatment with niacin–laropiprant was associated with a 55% increase in serious disturbances in diabetes control, most of which led to hospitalization (11.1% vs. 7.5%, P<0.001) (51). The extent to which the latter was due to laropiprant is unknown. Thus, care must be used in treating patients with diabetes with niacin. In patients in whom adjustments in diabetic therapy can easily be carried out the risk of adverse effects will likely be limited whereas in patients in whom adjustments in diabetic therapy will be difficult the risks of niacin therapy are likely to be increased. Careful patient selection and education are important steps to reduce the risks of niacin therapy in patients with diabetes.

 

Thus, while niacin therapy may adversely affect glucose homeostasis one needs to balance these adverse effects with the potential benefits of niacin therapy. One should note that in the Coronary Drug Project participants with abnormal glucose metabolism also demonstrated a decrease in cardiovascular events with niacin therapy (45).  

 

URIC ACID  

 

Niacin may increase uric acid levels by inhibiting the secretion of uric acid (8,77). In susceptible patients niacin therapy can precipitate gouty attacks (8).   

 

GASTROINTESTINAL SYMPTOMS  

 

Niacin therapy can induce heartburn, indigestion, nausea, diarrhea, and abdominal discomfort (8). High dose niacin is more likely to cause these gastrointestinal disturbances. The mechanism for these symptoms is not clear. 

 

MISCELLANEOUS  

 

Recent trials have reported an increased incidence of infections with niacin therapy (51,69). A trial of niacin in combination with laropiprant found increased bleeding (51). The increased bleeding could be due to the approximately 10% decrease in platelet levels that can occur with niacin (see Niaspan Package Insert). However, a very large observational study that compared rates of major gastrointestinal bleeding and intracranial hemorrhage in patients treated with niacin (>200,000 subjects) to propensity matched subjects on fenofibrate did not observe an increase in bleeding (78). Niacin has been reported to induce cystoid macular edema, which resolves when the drug is stopped (79).

 

Contraindications

 

There are a number of contraindications to niacin therapy (Table 7).

 

Table 7. Contraindications for Niacin Therapy

Active gastritis or peptic ulcer disease

Impaired liver function (elevated transaminases 2-3X the upper limit or cholestasis)

Uncontrolled gout

Pregnancy

Lactation

Poorly controlled diabetes

Active bleeding

 

Summary

 

The enthusiasm for the use of niacin has greatly decreased with the failure of AIM-HIGH and HPS-2 Thrive to show a decrease in cardiovascular events when niacin was added to statin therapy. In the absence of definitive data showing benefits from niacin therapy when added to a statin it is hard to justify the use of this drug given the frequent side effects. The availability of ezetimibe, bempedoic acid, and PCSK9 inhibitors has greatly reduced the need to use niacin to lower LDL-C levels. Additionally, in patients with markedly elevated TG levels (>500mg/dL), niacin can be employed in combination with other drugs to reduce the risk of pancreatitis but fibrates and omega-3-fatty acids are the initial choices.

 

OMEGA-3-FATTY ACIDS (FISH OIL)

 

Introduction

 

The lipid lowering effects of fish oil are mediated by two omega-3-fatty acids; eicosapentaenoic acid (C20:5n-3) (EPA) and docosahexaenoic acid (C22:6n-3) (DHA). There are four prescription products approved by the FDA which contain various amounts of EPA and DHA (Table 8). Lovaza and Omacor contain a mixture of EPA and DHA fatty acid esters (ethyl esters), Vascepa contains only EPA fatty acid esters (ethyl esters), and Epanova contains a mixture of EPA and DHA free fatty acids (Epanova is currently not available in the US).

 

Table 8. Prescription Omega-3-fatty acid products (data from package inserts)

Generic Name

Omega-3-ethyl esters

Icosapent ethyl

Omega-3-carboxylic acid

Brand Name

Lovaza or Omacor

Vascepa

Epanova

EPA/capsule

0.465g

1.0g

See below

DHA/capsule

0.375g

---

See below

Daily Dose

4 capsules/day

4 capsules/day

2-4 capsules/day

1-gram capsules of Epanova contain at least 850mg of fish oil derived fatty acids including multiple omega-3-fatty acids with EPA and DHA being the most abundant

 

Fish oil is also sold as a food supplement. It should be recognized that dietary fish oil supplements are not approved by the FDA and quality control will not meet the same rigorous standards as prescription or over the counter drugs. The amount of EPA and DHA can vary greatly in these supplements and one needs to read the labels carefully, as products can contain less than 100mg of EPA/DHA per 1 gram capsule (80). It is helpful to have the patient bring their fish oil supplements to the clinic for verification of the actual amount of EPA and DHA in the product. Moreover, the amount of EPA and DHA indicated on the label may not be accurate (81). One needs to take a sufficient number of capsules to provide 2-4 grams of EPA/DHA per day to effectively lower plasma TG levels. Depending upon the fish oil supplement, the patient may be required to take a large number of capsules to obtain 2-4 grams of EPA/DHA per day. Furthermore, these supplements may contain other compounds in addition to omega-3-fatty acids, such as cholesterol, oxidized lipids, and saturated fatty acids. The major advantage of fish oil supplements is that they are much less expensive than prescription omega-3-fatty acid drugs. If one elects to use fish oil supplements, one should have the patient use a single brand to try to ensure as much consistency as possible.

 

Some omega-3 supplements contain alpha linolenic acid (C18:3n-3) (ALA), a plant omega-3-fatty acid rather than EPA/DHA. ALA can be converted to EPA and DHA but the conversion is limited and hence it is ineffective in lowering plasma TG levels or altering other lipid or lipoprotein levels (82).

 

Effect of Omega-3-Fatty Acids on Lipid and Lipoprotein Levels

 

Table 9. Effect of Fish Oil Supplements on Lipids and Lipoproteins

Decreases TGs

No Change in Total Cholesterol

No Change in LDL-C; if TGs are very high may increase LDL-C

No Change in HDL-C

No Change in Lp(a); modest decrease in some studies

Shift from Small Dense LDL to Large Buoyant LDL

 

Several meta-analyses have examined the effect of fish oil supplements on lipid and lipoprotein levels. A meta-analysis by Eslick and colleagues of 47 studies with 16,511 participants found that fish oil supplements significantly decreased plasma TG levels by approximately 14% without resulting in clinically significant changes in total, LDL-C, or HDL-C levels (83). These authors also reported that the reduction in plasma TG levels was directly related to baseline plasma TG levels (i.e., the higher the baseline TG level the greater the reduction in TGs with fish oil). Additionally, the higher the dose of EPA/DHA, the greater the reduction in plasma TGs, with clinically significant reductions occurring with approximately 3.25 grams per day. A meta-analysis by Balk and colleagues of 21 studies also found minimal effects of fish oil supplements on total, LDL-C, and HDL-C levels (< 5% change) with significant decreases in plasma TG levels (most of the studies in this meta-analysis had at least a 15% decrease) (84). Similar to the meta-analysis by Eslick et al, the higher the baseline TG levels the greater the reduction in TG levels. 

 

Several meta-analyses have focused on specific patient populations. In a meta-analysis of patients with diabetes, twenty three trials with1075 participants were analyzed and similar to patients without diabetes the major effect of fish oil supplements was a reduction in plasma TG levels with no change in total cholesterol or HDL-C (85). A small increase in LDL-C was observed (4.3mg/dL). Of note, fish oil supplementation did not alter fasting glucose or glycated hemoglobin levels indicating that fish oil supplementation does not adversely affect glycemic control. In a meta-analysis that included patients with type 2 diabetes or impaired glucose metabolism a decrease in TGs was observed without significant changes in total cholesterol, LDL-C, or HDL-C levels (86). Again, no adverse effects on glycemic control were observed.

 

In patients with end stage renal disease several meta-analyses have consistently shown a decrease in plasma TGs with fish oil administration but the effect on total, LDL-C, and HDL-C has been variable (87-89). This variability was likely due to the small changes that were observed. In patients with nephrotic syndrome a study has shown a reduction in plasma TGs and an increase in LDL-C levels without a change in total cholesterol or HDL-C levels (90). In patients with non-alcoholic fatty liver disease, omega-3-fatty acids have also been shown to decrease plasma TG levels (91). Finally, In HIV infected subjects, fish oil supplementation was also effective in lowering plasma TG levels (92,93).

 

Thus, fish oil supplementation in a variety of different patient populations lowers plasma TG levels. In patients with elevated TG levels treated with 3-4 grams of EPA/DHA one can expect an approximate 25% decrease. Total plasma cholesterol levels are usually not altered by fish oil supplementation. The exceptions are patients with high chylomicron and/or VLDL levels where a substantial portion of the plasma cholesterol is carried on these TG rich lipoproteins. Fish oil supplementation will decrease the levels of these TG rich lipoproteins and thereby result in a decrease in total plasma cholesterol levels. LDL-C levels are not markedly affected by fish oil supplementation except in patients with very high TG levels (>500mg/dL) where increases in LDL-C levels have been observed (94-96). If there are sufficient reductions in plasma TG levels a shift from small dense LDL to large buoyant LDL may be observed (97,98). The effect of fish oil supplements on HDL-C levels is minimal except if the patient has very high TG levels where significant elevations (>10%) have been reported (94-96). Finally, some but not all studies have shown that the administration of fish oil modestly lowers Lp(a) levels (99-103)

 

During the development of pharmacological omega-3-fatty acid drugs for approval by the FDA, extensive clinical trials were carried out and will be reviewed below (Tables 10 and 11). It should be noted that these studies are not directly comparable as they studied different patient populations at different times.

 

EPA + DHA FATTY ACID ESTERS (LOVAZA)  

 

In patients with marked elevations in plasma TG levels (500-2000mg/dL) a 6 week trial of EPA + DHA esters resulted in a 31% decrease in plasma TGs, a 21% increase in LDL-C levels, and a 12% increase in HDL-C levels compared to the placebo group (96). In a 16 week trial TG concentrations were decreased by 45% and LDL-C and HDL-C were increased by 31% and 13%, respectively (94). Studies have also been carried out in patients with moderate hypertriglyceridemia (200-500mg/dL) who were on statin therapy (104). EPA + DHA esters resulted in a 23% decrease in plasma TGs and a 7% decrease in non-HDL-C levels, and a 4.6% increase in HDL-C levels (104).

 

EPA FATTY ACID ESTER ALONE (VASCEPA)  

 

In patients with marked elevations in plasma TGs (500-2000mg/dL), 4 grams of EPA ester alone significantly decreased TG levels by 33.1% and non-HDL-C levels by 17.7% (105). In contrast to EPA and DHA fatty acid esters, LDL-C and HDL-C levels were not significantly altered by EPA fatty acid esters alone (105). Studies have also been carried out in patients with moderate hypertriglyceridemia (200-500mg/dL) who were on statin therapy. EPA esters resulted in a 21.5% decrease in plasma TGs, 13.6% decrease in non-HDL-C, 6.2% decrease in LDL-C, and a 4.5% decrease in HDL-C levels (106)

 

EPA + DHA FATTY ACIDS (EPANOVA)  

 

In patients with marked elevations in plasma TGs (500-2000mg/dL), 4 grams of EPA + DHA fatty acids decreased plasma TGs by 31% and non-HDL-C by 9.6% and increased LDL-C by 19% and HDL-C by 5.8% (107). Studies have also been carried out in patients with moderate hypertriglyceridemia (200-500mg/dL) who were on statin therapy. EPA + DHA fatty acids resulted in a 20.6% decrease in plasma TGs, 6.9% decrease in non-HDL-C with no significant changes in LDL-C or HDL-C levels (95).

 

These studies demonstrate that in patients on statin therapy with moderate elevations in plasma TG levels the effects of these three pharmaceutical products on lipids and lipoprotein levels are similar (table 11). However, in patients with marked elevations in plasma TG levels EPA ethyl esters alone do not increase LDL-C levels whereas products containing EPA and DHA result in a substantial increase in LDL-C levels (table 10). It should also be noted that the ability of omega-3-fatty acids to reduce plasma TGs and increase HDL-C levels is enhanced if baseline TG levels are markedly elevated.

 

Table 10: Effect of Omega-3-Fatty Acids on Lipids and Lipoprotein in Patients with Marked Hypertriglyceridemia (500-2000mg/dL)

 

TGs

Non-HDL-C

LDL-C

HDL-C

EPA+DHA ethyl esters-

6 weeks

31% decrease

ND

21% increase

12% increase

EPA+DHA ethyl esters

12 weeks

45% decrease

ND

31% increase

13% increase

EPA ethyl esters

33% decrease

18% decrease

NS

NS

EPA+DHA fatty acids

31% decrease

9.6% decrease

19% increase

5.8% increase

ND- not determined; NS- no significant change

 

Table 11: Effect of Omega-3-Fatty Acids on Lipids and Lipoprotein in Patients with Moderate Hypertriglyceridemia (200-500mg/dL) on Statin Therapy

 

TGs

Non-HDL-C

LDL-C

HDL-C

EPA+DHA ethyl esters

23% decrease

7% decrease

__

4.6% increase

EPA ethyl esters

22% decrease

14% decrease

6.2% increase

4.5% decrease

EPA+DHA fatty acids

21% decrease

6.9% decrease

NS

NS

NS- no significant change

 

HEAD-TO-HEAD COMPARISONS  

 

A meta-analysis of six studies has compared the effect of EPA alone vs. DHA alone on plasma lipids and lipoproteins (108). Administration of DHA increased LDL-C by 4.6mg/dL compared to EPA (95% CI 2.2- 7.1). In contrast, DHA reduced plasma TG levels to a greater extent than EPA (6.1mg/dL; 95% CI 2.5- 9.8). Finally, DHA increased HDL-C levels more than EPA (3.7mg/dL; 95% CI: 2.4- 5.1). Whether these very modest differences are clinically significant is unknown.

 

Tatsuno et al compared the effect of DHA + EPA ethyl esters vs. EPA ethyl esters alone on lipid and lipoprotein levels in patients with mean baseline plasma TG of 250-270mg/dL and mean LDL-C levels of 125-135mg/dL (109,110). These authors found that at equivalent doses there were no differences in effect on plasma TG, LDL-C, or HDL-C levels between DHA + EPA ethyl ester or EPA ethyl ester treatment.

 

These head-to-head studies indicate that in subjects with moderate hypertriglyceridemia the effects of EPA and DHA on lipid and lipoprotein levels are similar. Perhaps if the baseline TGs were markedly elevated differences in response might have been observed.

 

IN COMBINATION WITH FENOFIBRATE  

 

In patients with marked hypertriglyceridemia a single drug is often not sufficient to lower TGs into the desired range. In patients with TG levels > 500mg/dL the combination of fenofibrate 130mg per day and 4 grams of Lovaza reduced TG levels to a greater extent than monotherapy with fenofibrate alone (7% decrease; P = 0.059) (111). Not unexpectedly, LDL-C levels were increased to a greater extent with combination therapy (9% increase; p= 0.03). When subjects who had received 8 weeks of fenofibrate monotherapy were treated with Lovaza during the 8-week, open-label extension study, TG levels were reduced by an additional 17.5% (P = 0.003). These results indicate that the addition of omega-3-fatty acids to fenofibrate will further decrease TG levels.

 

IN COMBINATION WITH NIACIN

 

Sixty patients with the metabolic syndrome were randomized to 16 weeks of treatment with placebo, Lovaza (4 g/day), extended release niacin (2 g/day), or both drugs in combination (17). In the niacin group TGs were decreased by 30%, in the omega-3-fatty acids group by 22%, and in the combination group by 42% compared to the placebo group. Of note, the beneficial effects of niacin on decreasing LDL and non-HDL-C were blunted by omega-3-fatty acids. These results show that the combination of niacin and fish oil will lower TG levels more than either drug individually but at the expense of diminishing the effect of niacin on LDL and non-HDL-C levels.

 

Mechanism Accounting for the Omega-3-Fatty Acid Induced Lipid Effects

 

As noted above, the major effect of fish oil is to lower plasma TG levels. The predominant cause of the reduction in plasma TG levels is a decrease in the hepatic production and secretion of TG rich lipoproteins (112-115). In cultured hepatocytes, omega-3-fatty acids inhibit the assembly and secretion of VLDL and apolipoprotein B 100 (113,115-117).  The incorporation of TGs into VLDL is a key regulatory step in determining the rate of formation and secretion of VLDL and there are a number of mechanisms by which omega-3 fatty acids reduce the level of hepatic TGs available for VLDL formation (112,113,115). Studies in animal models have demonstrated that omega-3-fatty acids inhibit fatty acid synthesis and stimulate fatty acid oxidation in the liver, which would reduce the availability of fatty acids for TG synthesis (112-115). The increase in fatty acid oxidation is due to omega-3-fatty acids activating PPAR alpha, which stimulates fatty acid oxidation in the liver and other tissues (112,114,115,118). The decrease in fatty acid synthesis is due to omega-3-fatty acids inhibiting the expression of SREBP-1c, a key transcription factor that regulates fatty acid synthesis (114,115,118). In addition, omega-3-fatty acids decrease TG synthesis, which may be due to the decreased availability of fatty acids and an inhibition of the activity of DGAT, a key enzyme required for TG synthesis (112,114,115). Finally, omega-3-fatty acids also decrease the flux of free fatty acids from adipose tissue to the liver, which will lead to a decreased quantity of fatty acids available for TG synthesis in the liver (112). This decrease in flux of free fatty acids is due to omega-3-fatty acids reducing hormone sensitive lipase mediated intracellular lipolysis in adipose tissue (112). It is likely that these and perhaps other factors lead to the decreased availability of TGs resulting in a reduction in VLDL formation and secretion. In addition, the peroxidation of omega-3-fatty acids may stimulate the degradation of apolipoprotein B-100, which would provide another pathway that could contribute to a decrease in VLDL formation and secretion (115).

 

While not the primary mechanism for the decrease in plasma TGs, studies have shown that omega-3-fatty acids may increase the clearance of TG rich lipoproteins (112,119). Post heparin lipoprotein lipase activity is not increased by omega-3-fatty acid administration but the lipolytic activity of non-stimulated plasma is enhanced (112,119).  Additionally, apolipoprotein C-III levels are decreased with omega-3-fatty acid administration which could also contribute to an increase in the clearance of TG rich lipoproteins (120-123).

 

The increase in LDL-C levels that occurs in patients with marked hypertriglyceridemia treated with omega-3-fatty acids is thought to be due to the enhanced conversion of VLDL to LDL (114). The increase in HDL-C observed in studies in patients with very high TG levels may be due to the increased clearance of TG rich lipoproteins.    

 

Pharmacokinetics and Drug Interactions

 

Omega-3 ethyl esters and fatty acids are absorbed by the GI tract similar to other dietary lipids. It is worth noting that omega-3-free fatty acids (Epanova) are directly absorbed by the small intestine and are not dependent on pancreatic lipases for absorption. Thus, absorption of omega-3-fatty acids is not decreased in patients with pancreatic insufficiency and therefore may be preferred in patients with pancreatic disease. Additionally, the bioavailability of omega-3-fatty acids with a low fat diet was greater than omega-3-ethyl esters while there was little difference between these different formulations with a high fat diet (124,125).

 

Drug interactions have not been seen with omega-3-fatty acids (Package Inserts for Lovaza, Vascepa, and Epanova).

 

Effect of Low Dose Omega-3-Fatty Acids on Clinical Outcomes

 

Initial studies of the effect of low dose fish oil administration on cardiovascular outcomes were favorable, demonstrating a reduction in events including all-cause mortality. However, more recent studies have failed to confirm these favorable results. In these more recent studies the use of other drugs, such as statins, that reduce cardiovascular disease were more intensively utilized. The outcomes studies that will be described below were carried out with doses of EPA and DHA that are lower than the doses used to lower plasma TGs. We will limit our discussion to the administration of fish oil as a drug and not discuss diet studies, such as DART, which had patients increase fatty fish intake (126,127).

 

  • GISSI-Prevenzione trial was a randomized trial of 850-882mg of EPA and DHA ethyl esters per day in 11,323 participants with a recent myocardial infarction (< 3 months) for 3.5 years (128). The primary endpoint was death, non-fatal myocardial infarction, and stroke. No change in total cholesterol, LDL-C, or HDL-C was observed but plasma TG levels were decreased by 5%. Patients treated with EPA/DHA had a significant decreased risk of major cardiovascular events (RR 0.90), cardiac death (RR 0.78), and sudden death (RR 0.74). The decrease in sudden death occurred very quickly and was noted as early as 4 months after initiation of therapy. Interestingly, non-fatal cardiovascular events were not affected by EPA/DHA treatment (RR 0.98). The decrease in total mortality was driven by a reduction in sudden death suggesting an anti-arrhythmic effect of EPA/DHA.

 

  • GISSI-Heart Failure (GISSI-HF) trial was a randomized, double-blind, placebo-controlled trial in patients with chronic heart failure who were randomly assigned to 850-882mg of EPA and DHA ethyl esters per day (n=3,494) or placebo (n=3,481) (129). Patients were followed for a median of 3.9 years. Primary endpoints were time to death, and time to death or admission to the hospital for cardiovascular reasons. Omega-3-fatty acid treatment at these low doses resulted in a slight decrease in plasma TG levels with no change in total, LDL-C or HDL-C levels. In the omega-3-fatty acid group 27% patients died from any cause vs. 29% in the placebo group (HR 0.91; p=0.041). In the omega-3-fatty acid group 57% of patients died or were admitted to hospital for cardiovascular reasons vs. 59% in the placebo group (HR 0.92; p=0.009). No significant differences were observed in fatal or non-fatal myocardial infarctions or strokes. In this trial, similar to the GISSI-Prevenzione trial, the benefit was primarily due to a reduction in arrhythmic events and little benefit on atherothrombotic events was noted.

 

  • OMEGA was a randomized, placebo-controlled, double-blind, trial in 3,851 survivors of an acute myocardial infarction (130). Patients were randomized 3 to 14 days after an acute myocardial infarction to omega-3-acid ethyl esters, 1 gram/day (460mg EPA and 380mg DHA) or placebo capsules containing 1 gram of olive oil and followed for one year. The primary endpoint was rate of sudden death and secondary end points were total mortality and nonfatal clinical events. No significant differences were seen in the primary or secondary endpoints.

 

  • Alpha Omega was a double-blind, placebo-controlled trial in 4,837 patients between 60 and 80 years of age (78% men) who had had a myocardial infarction (131). Patients were randomized to receive for 40 months one of four trial margarines: a margarine supplemented with a combination of EPA and DHA (with a targeted additional daily intake of 400mg of EPA-DHA; actual intake 226mg EPA and 150mg DHA), a margarine supplemented with alpha-linolenic acid (ALA) (with a targeted additional daily intake of 2g of ALA), a margarine supplemented with EPA-DHA and ALA, or a placebo margarine. The primary end point was the rate of major cardiovascular events, which comprised fatal and nonfatal cardiovascular events and cardiac interventions. Neither low dose EPA-DHA, ALA, nor the combination of EPA/DHA and ALA significantly reduced the rate of major cardiovascular events or cardiac interventions.

 

  • FOL.OM3 Study was a double blind, randomized, placebo-controlled trial in 2,501 patients with a history of a myocardial infarction, unstable angina, or ischemic stroke in the past 12 months (132). Patients were randomized to a daily dietary supplement containing 5-methyltetrahydrofolate (560μg), vitamin B-6 (3mg), and vitamin B-12 (20μg) or placebo; and a dietary supplement containing omega 3 fatty acids (600mg of EPA and DHA) or placebo. Median duration of treatment was 4.7 years. The primary outcome was a composite of non-fatal myocardial infarction, stroke, or death from cardiovascular disease. Treatment with B vitamins or omega 3 fatty acids had no significant effect on major vascular events.

 

  • Origin was a double-blind study in 12,536 patients at high risk for cardiovascular disease who had impaired fasting glucose, impaired glucose tolerance, or diabetes (133). Patients were randomized to receive a 1-gram capsule containing at least 900mg of ethyl esters of omega-3 fatty acids (EPA 465mg and DHA 375mg) or placebo for approximately 6 years. The primary outcome was death from cardiovascular causes. TG levels were reduced by 14.5mg/dL in the group receiving omega-3-fatty acids compared to the placebo group (P<0.001), without a significant effect on other lipids. The incidence of the primary outcome was not significantly decreased among patients receiving omega-3-fatty acids as compared with those receiving placebo. The use of omega-3-fatty acids also had no significant effect on the rates of major vascular events, death from any cause, or death from arrhythmia.

 

  • Risk and Prevention Study was a double-blind, placebo-controlled trial in 12,513 men and women with multiple cardiovascular risk factors or atherosclerotic vascular disease but not myocardial infarctions (134). Patients were randomly assigned to 1-gram daily omega-3 fatty acids (EPA and DHA content not <85 %,) or placebo (olive oil) for 5 years. The initially specified primary end point was the rate of death, nonfatal myocardial infarction, and nonfatal stroke. At 1 year, after the event rate was found to be lower than anticipated, the primary end point was revised as time to death from cardiovascular causes or admission to the hospital for cardiovascular causes. Plasma TG levels decreased slightly more in the omega−3-fatty acid group than in those who received placebo (−28.2±1.3mg/dL vs. −20.1±1.3mg/dL; P<0.001). Total, LDL, and HDL-C levels were similar in the omega-3-fatty acid and placebo groups. No significant differences were observed between the omega-3-fatty acid group and placebo group for the primary endpoint or any of the secondary endpoints.

 

  • A Study of Cardiovascular Events in Diabetes (ASCEND) was a randomized, placebo controlled, double blind trial of 1-gram omega-3-fattys acids (400mg EPA and 300mg DHA ethyl esters) vs. olive oil placebo in 15,480 patients with diabetes without a history of cardiovascular disease (primary prevention trial) (135). The primary end point was serious vascular events (non-fatal myocardial infarction, non-fatal stroke, transient ischemic attack, or vascular death). Total cholesterol, HDL-C, and non-HDL-C levels were not significantly altered by omega-3-fatty acid treatment (changes in TG levels were not reported). After a mean follow-up of 7.4 years the composite outcome of a serious vascular event or revascularization occurred in 882 patients (11.4%) on omega-3-fatty acids and 887 patients (11.5%) on placebo (rate ratio, 1.00; 95% CI, 0.91 to 1.09). Serious adverse events were similar in placebo and omega-3-fatty acid treated groups.

 

  • The Vitamin D and Omega-3 Trial (Vital) was a randomized, double blind, placebo-controlled trial of 1-gram omega-3 fatty acids (465mg EPA and 375mg DHA ethyl esters) vs. placebo in 25,875 men (>50 years of age) and women (>55 years of age) that were not selected on the basis of an elevated risk (primary prevention) (136). Changes in lipid levels were not reported. The primary end point was major cardiovascular events, a composite of myocardial infarction, stroke, or death from cardiovascular causes. After a median follow-up of 5.3 years, major cardiovascular event occurred in 386 participants in the omega-3 fatty acid group and in 419 in the placebo group (hazard ratio, 0.92; 95% confidence interval (CI), 0.80 to 1.06; P=0.24). Serious adverse events were similar in placebo and omega-3-fatty acid treated groups.

 

  • Summary: The above results indicate that low dose fish oil (doses that do not greatly affect lipid levels) do not consistently reduce the risk of cardiovascular disease.

 

Effect of High Dose Omega-3-Fatty Acids on Clinical Outcomes

 

  • Japan EPA Lipid Intervention Study (JELIS) was an open label study without a placebo in patients with total cholesterol levels > 254mg/dL with (n= 3,664) or without cardiovascular disease (n=14,981) who were randomly assigned to be treated with 1800 mg of EPA (Vascepa) + statin (n=9,326) or statin alone (n= 9,319) with a 5-year follow-up (130). The primary endpoint was any major coronary event, including sudden cardiac death, fatal and non-fatal myocardial infarction, and other non-fatal events including unstable angina pectoris, angioplasty, stenting, or coronary artery bypass grafting. Total, LDL-C, and HDL-C levels were similar in the two groups but plasma TGs were modestly decreased in the EPA treated group (5% decrease in EPA group compared to controls; p = 0.0001). In the EPA group the primary endpoint occurred in 2.8% of the patients vs. 3.5% of the patients in the statin alone group (19% decrease; p = 0.011). Unstable angina and non-fatal coronary events were also significantly reduced in the EPA group but in this study sudden cardiac death and coronary death did not differ between groups. Unstable angina was the main component contributing to the primary endpoint and this is a more subjective endpoint than other endpoints such as a myocardial infarction, stroke, or cardiovascular death. In patients with high TG levels (>150 mg/dL) and low HDL-C levels (<40 mg/dL EPA treatment decreased the risk of CAD by 53% (HR: 0.47; P=0.043) (137). A subjective endpoint has the potential to be an unreliable endpoint in an open label study and is a limitation of the JELIS Study.

 

  • The Reduction of Cardiovascular Events with EPA – Intervention Trial (REDUCE-IT) was a randomized, double blind trial of 2 grams twice per day of EPA ethyl ester (icosapent ethyl) (Vascepa) vs. mineral oil placebo in 8,179 patients with hypertriglyceridemia (135mg/dL to 499mg/dL) and established cardiovascular disease or high cardiovascular disease risk (diabetes plus one risk factor) who were on stable statin therapy (138). The primary end point was a composite of cardiovascular death, nonfatal myocardial infarction, nonfatal stroke, coronary revascularization, or unstable angina. The key secondary end point was a composite of cardiovascular death, nonfatal myocardial infarction, or nonfatal stroke. At baseline, the median LDL-C level was 75.0 mg/dL, HDL-C level was 40.0 mg/dL, and TG level was 216.0 mg/dL. The median change in TG level from baseline to 1 year was a decrease of 18.3% (−39.0 mg/dL) in the EPA group and an increase of 2.2% (4.5 mg/dL) in the placebo group. After a median of 4.9 years the primary end-point occurred in 17.2% of the patients in the EPA group vs. 22.0% of the patients in the placebo group (hazard ratio, 0.75; P<0.001), indicating a 25% decrease in events. The number needed to treat to avoid one primary end-point event was 21. The reduction in cardiovascular events was noted after approximately 2 years of EPA treatment. Additionally, the rate of cardiovascular death was decreased by 20% in the EPA group (4.3% vs. 5.2%; hazard ratio, 0.80; P=0.03). The cardiovascular benefits of EPA were similar across baseline levels of TGs (<150, ≥150 to <200, and ≥200 mg per deciliter). Moreover, the cardiovascular benefits of EPA appeared to occur irrespective of the attained TG level at 1 year (≥150 or <150 mg/dL), suggesting that the cardiovascular risk reduction was not associated with attainment of a normal TG level. An increase in hospitalization for atrial fibrillation or flutter (3.1% vs. 2.1%, P=0.004) occurred in the EPA group. In addition, serious bleeding events occurred in 2.7% of the patients in the EPA group and in 2.1% in the placebo group (P=0.06). There were no fatal bleeding events in either group and the rates of hemorrhagic stroke, serious central nervous system bleeding, and serious gastrointestinal bleeding were not significantly higher in the EPA group than in the placebo group.

 

It should be noted that in this trial mineral oil was used as the placebo. In the placebo group the LDL-C, non-HDL-C, and CRP levels were increased compared to the EPA group during the trial (LDL-C 96mg/dL vs 85mg/dL; non-HDL-C 130mg/dL vs. 113mg/dL; hsCRP 2.8mg/L vs. 1.8mg/L). The impact of these adverse changes on clinical outcomes is uncertain and whether they contributed to the apparent beneficial effects observed in the individuals treated with EPA is unknown.

 

  • The STRENGTH Trial was a double-blind, randomized, trial comparing 4 grams per day of a carboxylic acid formulation of omega-3 fatty acids (EPA and DHA; Epanova) (n = 6,539)) vs. corn oil placebo (n = 6539) in statin-treated participants with high cardiovascular risk, hypertriglyceridemia, and low levels of HDL-C (139). Approximately 55% of patients had established cardiovascular disease and approximately 70% had diabetes. Median LDL-C level was 75.0 mg/dL, median TG level was 240 mg/dL and median HDL-C level was 36 mg/dL. There were minimal differences in the change in LDL-C and HDL-C levels between the treated and placebo groups after treatment for 12 months but as expected there was a greater reduction in TG levels in the group treated with omega-3-fatty acids (−19.0% vs −0.9%). The primary endpoint was a composite of cardiovascular death, nonfatal myocardial infarction, nonfatal stroke, coronary revascularization, or unstable angina requiring hospitalization which occurred in 12.0% of individuals treated with omega-3 CA vs. 12.2% treated with corn oil (hazard ratio, 0.99; P = .84). There were no significant differences between the treatment groups with regard to the risk of the individual components of the primary end point over the 3-4 years of the study. Similar to the REDUCE-IT trial atrial fibrillation was increased with EPA + DHA treatment (HR 1.69 CI 1.29- 2.21). Thus, in contrast the JELIS and REDUCE-IT trials the STRENGTH trial did not demonstrate a benefit of treatment with a mixture of omega-3-fatty acids (EPA + DHA).

 

  • The OMEMI trial was a randomized trial of 1.8 grams per day of omega-3-fatty acids (930 mg EPA and 660 mg DHA) (n= 505) vs. corn oil placebo (509) in patients aged 70 to 82 years with a recent myocardial infarction (2-8 weeks) (140). Baseline LDL-C was approximately 76mg/dL, HDL-C was 49mg/dL, and TGs 110mg/dL. The primary endpoint was a composite of nonfatal myocardial infarction, unscheduled revascularization, stroke, all-cause death, and heart failure hospitalization after 2 years of follow-up. The primary endpoint occurred in 21.4% of patients on omega-3-fatty acids vs. 20.0% on placebo (hazard ratio, 1.08; P=0.60). TGs levels decreased 8.1% in the omega-3-fatty acid group and increased 5.1% in the placebo group (between group difference 13.2%; P<0.001) while changes in LDL-C were minimal in both groups. Thus, similar to the STRENGTH trial no benefits on cardiovascular disease were observed with EPA + DHA treatment.

 

Summary of Omega-3-Fatty Acid Clinical Outcome Trials

 

  • Low dose omega-3-fatty acids are not effective at decreasing cardiovascular outcomes.
  • High dose EPA (JELIS and REDUCE-IT) reduced cardiovascular outcomes while high dose EPA+DHA (STENGTH and OMEMI) did not decrease cardiovascular outcomes.
  • The decrease in TG levels is not a major contributor to the beneficial effect of high dose EPA as the combination of high EPA+DHA lowers TG levels to the same degree as EPA alone without benefit. Additionally, the JELIS trial only lowered TG levels by 5% but nevertheless reduced cardiovascular events. It is likely that the beneficial effects of EPA seen in the JELIS and REDUCE-IT trials are multifactorial with TG lowering making only a small contribution to the decrease in cardiovascular disease. Other actions of EPA, such as decreasing platelet function, anti-inflammation, decreasing lipid oxidation, stabilizing membranes, etc. could account for or contribute to the reduction in cardiovascular events (141). A large meta-analysis, excluding the REDUCE-IT trial, demonstrated that a 40mg/dL decrease in triglyceride levels resulted in a relative risk reduction of only 0.96 (4% decrease) indicating that one needs to markedly lower triglyceride levels to reduce cardiovascular events (142).
  • Whether EPA has special properties that resulted in the reduction in cardiovascular events in the REDUCE-IT trial or there were flaws in the trial design (the use of mineral oil as the placebo) is uncertain and debated. It should be noted that in the REDUCE-IT trial LDL-C and non-HDL-C levels were increased by approximately 10% in the mineral oil placebo group (138). Additionally, Apo B levels were increased by 7% (6mg/dL) by mineral oil (138). Finally, an increase in hsCRP (20-30%) and other biomarkers of atherosclerosis (oxidized LDL-C, IL-6, IL-1 beta, and lipoprotein-associated phospholipase A2) were noted in the mineral oil group (138,143). In the STRENGTH trial there were no differences in LDL-C, Non-HDL-C, HDL-C, Apo B, or hsCRP levels between the treated vs. placebo groups (139). Whether EPA has special properties compared to DHA leading to a reduction in cardiovascular events or the mineral oil placebo resulted in adverse changes increasing ASCVD in the placebo resulting in an artifactual decrease in the EPA group is debated (144,145). Ideally, another large randomized cardiovascular trial with EPA ethyl ester (icosapent ethyl) (Vascepa) using a placebo other than mineral oil would help resolve this controversy. In the meantime, clinicians will need to use their clinical judgement on whether to treat patients with modest elevations in TG levels with EPA (icosapent ethyl; Vascepa) balancing the potential benefits of treatment vs. the potential side effects.

 

Side Effects

 

Gastrointestinal side effects such as diarrhea, nausea, dyspepsia, abdominal discomfort, and eructation have been observed with fish oil therapy (Package Inserts for Lovaza, Vascepa, and Epanova).

 

At very high doses, omega-3-fatty acids can inhibit platelets and prolong bleeding time. However, at the recommended doses this has not been a major clinical problem but nevertheless when patients are on anti-platelet drugs one should be alert for the possibility of bleeding problems (Package Inserts for Lovaza, Vascepa, and Epanova). Increased bleeding was noted in the REDUCE-IT trial in the patients treated with icosapent ethyl 4 grams/day (EPA) (see above discussion of this trial). A recent review found no evidence for discontinuing the use of omega-3 fatty acid treatment before invasive procedures or when given in combination with other agents that affect bleeding (146).

 

As noted above an increase in atrial fibrillation was observed in the REDUCE-IT trial in the patients treated with icosapent ethyl 4 grams/day (EPA) and in the STRENGTH trial in the patients treated with EPA + DHA.

 

Contraindications

 

There are no contraindications to the use of omega-3-fatty acids. Lovaza, Omacor, and Vascepa are pregnancy category C drugs and they should only be used if the benefits to the mother outweigh the potential risks to the fetus.

 

Conclusions

 

Omega-3-fatty acids are effective drugs in reducing TG levels with few significant side effects, drug interactions, or contraindications.  High dose EPA (4 grams/day) reduced cardiovascular disease events in the REDUCE-IT trial and a moderate dose of EPA (1.8 grams/day) reduced cardiovascular events in the JELIS trial but trials of EPA and DHA have not produced cardiovascular benefits. The basis for these differences is debated and discussed in the “Summary of Omega-3-Fatty Acid Clinical Outcome Trials” section above. Finally, omega-3-fatty acids are effective in lowering TGs in patients with marked hypertriglyceridemia and while not proven will likely reduce the risk of development of pancreatitis.

 

FIBRATES

 

Introduction

 

The fibrate drug class includes clofibrate, gemfibrozil, fenofibrate, bezafibrate, and ciprofibrate. Clofibrate was developed in the 1960s and was the first member of this class. Clofibrate is no longer available because of an increased risk of adverse effects. Gemfibrozil and fenofibrate are available in the United States while gemfibrozil, fenofibrate, bezafibrate, and ciprofibrate are available in Europe. All of the fibrates work via activation of the nuclear hormone receptor PPAR alpha.

 

Effect of Fibrates on Lipid and Lipoprotein Levels

 

Table 12. Effect of Fibrates on Lipids and Lipoproteins

Decreases TG

Increases HDL-C

Decreases LDL-C; if TGs Very High can Increase LDL-C

Decreases Non-HDL-C

Decreases Apolipoprotein B

Decreases LDL Particle Number

Shift Small Dense LDL to Large Buoyant LDL

No Effect on Lp(a)

 

Fibrates reduce fasting TG levels by 25-50% (147-149). The magnitude of the reduction in TGs is dependent on the baseline TG levels. Patients with marked elevations in TGs have a greater reduction in TG levels (147,149,150). In addition, fibrates increase HDL-C levels by 5-20% (148,149). The increase in HDL-C levels is more robust if the TG levels are elevated and/or if the HDL-C levels are low (150). The effect on LDL-C is more variable (149). If the TG levels are very high (>400-500mg/dL), fibrate therapy may result in an increase in LDL-C levels whereas if TGs are not elevated fibrates decrease LDL-C by 10-30% (147). Given the decrease in plasma TGs and LDL-C levels, fibrates also reduce apolipoprotein B, LDL particle number, and non-HDL-C levels (149). Depending upon the TG level there may be a shift from small dense LDL towards large LDL particles (149). Fibrates do not have any major or consistent effects on Lp(a) levels (151). Table 13 below shows the effect of fenofibrate on lipid and lipoprotein levels in patients with different lipid profiles and illustrates some of the principles outlined above.

 

Table 13. Effect of Fenofibrate on Lipid and Lipoprotein Levels

 

TGs

LDL-C

HDL-C

Elevated TG Levels

 

 

 

Baseline Levels

~404mg/dL

~125mg/dL

~35mg/dL

Change with Fenofibrate

45% Decrease

2.5% Increase

16% Increase

Elevated LDL-C and TG Levels

 

 

 

Baseline Levels

232mg/dL

220mg/dL

46.7mg/dL

Change with Fenofibrate

37% Decrease

13% Decrease

12% Increase

Elevated LDL-C and Normal TG Levels

 

 

 

Baseline Levels

102mg/dL

228mg/dL

58.1mg/dL

Change with Fenofibrate

35% Decrease

29% Decrease

7% Increase

The values are adjusted for changes in the placebo group. Data modified from Tricor Package Insert.

 

In large, randomized, fibrate outcome trials similar changes in lipid and lipoprotein levels were noted (Table 14). These trials are discussed in detail in the effect of fibrates on cardiovascular outcomes section presented below.

 

Table 14. Effect of Fibrates on Lipid and Lipoprotein Levels in Large Outcome Studies*

 

TGs

LDL-C

HDL-C

Helsinki Heart Study- Gemfibrozil (152)

35% Decrease

11% Decrease

10% Increase

VA-HIT Study

Gemfibrozil (153)

31% Decrease

No Change

6% Increase

BIP Study

Bezafibrate (154)

21% Decrease

7% Decrease

18% Increase

Leader Study

Bezafibrate (155)

23% Decrease

8% Decrease

8% Increase

Field Study

Fenofibrate (156)

29% Decrease

12% Decrease

5% Increase

*The values are adjusted for changes in the placebo group.

 

The different fibrates in general cause similar changes in lipid and lipoprotein levels. There are only a few comparative trials of fibrates comparing their effects on lipid and lipoprotein levels and these trials have been very small. Comparisons of ciprofibrate and gemfibrozil have not shown any major differences between these two fibrates (157,158). In contrast, two very small trials have compared gemfibrozil vs. fenofibrate and reported that fenofibrate was more efficacious in lowering LDL levels than gemfibrozil (159,160).

 

In very rare instances fibrates can cause a paradoxical marked decrease in HDL-C levels (161-164). This rare paradoxical decrease in HDL-C typically occurs when fibrates are used in combination with a thiazolidinedione (rosiglitazone and pioglitazone) but can occur when fibrates are used alone or with ezetimibe (161-165). The decrease in HDL-C can be extreme with decreases of 50% to 88% reported and recovery to normal can take weeks after the fibrate is discontinued (162). The mechanism for this paradoxical effect is unknown.

 

Effect of Fibrates in Combination with Other Lipid Lowering Drugs on Lipid and Lipoprotein Levels

 

STATINS

 

Statins are the primary drugs used to treat most patients with dyslipidemia. Statins are very effective in lowering LDL-C levels but have only modest effects on TG and HDL-C levels. Therefore, it is appealing to add a fibrate to patients who on statin therapy have LDL-C levels at goal but still have elevated non-HDL-C and TG levels and decreased HDL-C levels. Therefore, there have been numerous studies examining the effect of the combination of statins and fibrates on lipid and lipoprotein levels. An example is the Safari Trial which compared the effect of simvastatin only (n=207) vs. simvastatin + fenofibrate (n=411) in patients with combined hyperlipidemia (166). The results of this trial are shown in table 15. As anticipated, adding a fibrate results in a further lowering of LDL-C, non-HDL-C, and TG levels with a further increase in HDL-C.

 

Table 15. Effect of Simvastatin Alone vs. Simvastatin + Fenofibrate on Lipid and Lipoprotein Levels

 

LDL

TG

Non-HDLC

HDL

Simvastatin

-26%

-20%

-26%

+10%

Simvastatin + Fenofibrate

-31%

-43%

-35%

+19%

 

A meta-analysis of 9 studies with over 1,200 participants compared the effect of statin alone vs. statin + fibrate on lipid and lipoprotein levels (167). The combination of statins and fibrates provided significantly greater reductions in total cholesterol, LDL-C, and TGs, and a significantly greater increase in HDL-C than treatment with statins alone. A larger meta-analysis of 13 randomized controlled trials, involving 7,712 patients, similarly demonstrated significant decreases in LDL-C (8.8mg/dL), TGs (58mg/dL), and total cholesterol (11.2mg/dL), and increases in HDL-C (4.65mg/dL) in patients receiving the combination of statins + fibrates compared with statin therapy alone (168). The combination of statins + fibrates also result in a shift of LDL particles from small dense particles to large buoyant particles whereas no change in LDL particle size was observed with statin monotherapy (169).  

 

A recent meta-analysis of 6 studies with over 400 participants compared the effect of adding a statin to fibrate therapy (fibrate alone vs. fibrate + statin) and showed similar changes (170).  The fibrate-statin combination produced significantly greater reductions in the levels of total cholesterol, LDL-C, and TGs compared to fibrate alone. In contrast there was no significant difference in HDL-C levels in the fibrate vs. fibrate + statins group.

 

EZETIMIBE

 

In patients unable to tolerate statin therapy one needs to use other drugs to treat dyslipidemia. In a study comparing the effect of ezetimibe 10mg alone, fenofibrate 145mg alone, or ezetimibe + fenofibrate the combination had a better effect on the lipid profile resulting in a greater decrease in LDL-C levels and increase in HDL-C levels than either drug alone (Table 16) (171).

 

Table 16. Effect of the Combination of Ezetimibe and Fenofibrate on Lipid and Lipoprotein Levels

 

LDL-C

HDL-C

TG

Ezetimibe

23% Decrease

2.2% Increase

10% Decrease

Fenofibrate

22% Decrease

7.5% Increase

38% Decrease

Ezetimibe + Fenofibrate

34% Decrease

11.5% Increase

38% Decrease

 

Similar results were observed in another randomized trial of ezetimibe 10mg and fenofibrate 160mg (172). Moreover, both fibrate therapy and the combination of ezetimibe and fenofibrate results in a shift of LDL particles from small dense LDL particles to large buoyant particles (172).

 

EZETIMIBE + STATIN

 

A large randomized trial has compared the effect of ezetimibe /simvastatin 10mg/20mg, fenofibrate 160mg, or ezetimibe/simvastatin + fenofibrate on lipid and lipoprotein levels. As one would expect triple drug therapy had a better effect on the lipid profile (Table 17) (173). While ezetimibe/simvastatin was very effective in lowering LDL-C levels and fenofibrate was very effective in lowering TGs and raising HDL-C levels the combination resulted in more favorable changes in TGs. In a similar study the addition of fenofibrate 135mg to atorvastatin 40 mg + ezetimibe 10 mg resulted in a greater reduction in TGs (-57% vs. -40%; p<0.001) and a greater increase in HDL (13% vs. 4.2%; p<0.001) than placebo (174).  Fibrate therapy and ezetimibe/simvastatin + fenofibrate also resulted in a shift of LDL particles from small dense LDL particles to large buoyant particles (173).

 

Table 17. Effect of the Combination of Ezetimibe/Simvastatin and Fenofibrate on Lipid and Lipoprotein Levels

 

LDL-C

HDL-C

TG

Placebo

-3.5%

+1.1

-3.1%

Ezetimibe/Simvastatin

-47%

+9.3%

-29%

Fenofibrate

-16%

+18.2

-41

Eze/Simva + Fenofibrate

-46%

+18.7

-50%

 

BILE ACID SEQUESTRANT  

 

Studies have also examined the effect of fibrates in combination with bile acid sequestrants. Participants receiving fenofibrate 160 mg/day were randomized to receive either colesevelam HCl 3.75 g/day or placebo (175). No significant differences in TG or HDL-C levels were observed between the two groups. However, LDL-C levels were decreased in the fenofibrate + colesevelam group compared to the fenofibrate + placebo group (12.4% greater decrease: p<0.001). A study of the combination of fenofibrate and colestipol also demonstrated a more marked decrease in LDL-C with that combination compared to either drug alone (colestipol -18%; fenofibrate -17%, colestipol + fenofibrate 37%) (176). The combination of both drugs did not blunt the effects of fenofibrate on VLDL and HDL. Other studies of the combination of a fibrate with a bile acid sequestrant have also demonstrated an enhanced effect in lowering LDL-C levels (177-179).

 

NIACIN

 

Surprisingly there are few large randomized trials examining the effect of combination therapy with niacin + fibrate vs. monotherapy. One very small trial did reported that while both niacin monotherapy and bezafibrate monotherapy were effective in lowering serum TGs there was no added benefit of combination therapy in reducing serum TG level although a large variance may have reduced the ability to detect statistically significant results (16). A larger trial in HIV+ patients reported that the combination of niacin and fenofibrate was better at lowering TGs and non-HDL-C and increasing HDL-C levels than monotherapy with either niacin or fenofibrate (17). It would be informative if additional trials of fibrate + niacin combination therapy were carried out in patients with marked hypertriglyceridemia that can often be difficult to control with lifestyle changes and monotherapy.

 

FISH OIL  

 

In patients with TG levels > 500mg/dL the combination of fenofibrate 130mg per day and 4 grams of Lovaza (DHA and EPA) reduced TG levels to a greater extent than monotherapy with fenofibrate alone (7% decrease; P = 0.059) (103). Not unexpectedly, LDL levels were increased to a greater extent with combination therapy (9% increase; p= 0.03). When subjects who had received 8 weeks of fenofibrate monotherapy were treated with Lovaza (DHA and EPA) during the 8-week, open-label extension study TG levels were reduced by an additional 17.5% (P = 0.003). These results indicate that the addition of omega-3-fatty acids to fenofibrate will further decrease TG levels.

 

Mechanisms Accounting for the Fibrate Induced Lipid Effects

 

Fibrates are ligands that bind and activate PPAR alpha, a member of the family of nuclear hormone receptors that are activated by lipids (180,181). PPAR alpha is highly expressed in the liver and other tissues important in fatty acid metabolism. PPAR alpha forms a heterodimer with RXR and together the PPAR alpha:RXR complex when activated binds to the PPAR response elements in a large number of genes and regulates the expression of these genes (180,181). The natural ligands of PPAR alpha are fatty acid derivatives formed during lipolysis, lipogenesis, or fatty acid catabolism (180,181).

 

TRIGLYCERIDES  

 

Fibrates lower plasma TG levels by decreasing VLDL production and by increasing the clearance of TG rich lipoproteins (182,183). The decrease in VLDL production is primarily due to PPAR alpha activation of the beta oxidation of fatty acids, which reduces the substrate available for the synthesis of TGs and the formation of VLDL (180,183). Additionally, a decrease in hepatic fatty acid synthesis may also contribute to the decrease in fatty acids (180,183). The increased clearance of TG rich lipoproteins is due to PPAR alpha stimulating the transcription of lipoprotein lipase, the key enzyme that catabolizes the TGs carried by VLDL and chylomicrons (180,183). In addition, activation of PPAR alpha also inhibits the transcription of APO C-III, which inhibits lipoprotein lipase activity (180,183). A decrease in Apo C-III enhances the clearance of TG rich lipoproteins by increasing lipoprotein lipase activity. Notably, a decrease in Apo C-III also decreases TG levels in patients deficient in lipoprotein lipase indicating that there are multiple mechanisms for its effects on TG metabolism (184). Recent studies suggest that Apo C-III inhibits the uptake of TG rich lipoproteins into the liver by the LDL receptors/ LDLR-related protein 1 axis (185). PPAR alpha activation also increases the transcription of Apo A-V, which would also facilitate the activity of lipoprotein lipase (180).

 

HIGH DENSITY LIPOPROTEINS

 

The increase in HDL induced by fibrates is due to PPAR alpha activation stimulating Apo A-I and A-II transcription (180,183). This leads to the increased production of HDL (182). In addition, a decrease in TG rich lipoproteins may result in a reduction in CETP mediated transfer of cholesterol from HDL to VLDL and of TG from VLDL to HDL (183). This would lead to less TG enrichment of HDL and a decrease in the opportunity of hepatic lipase to remove TG leading to small HDL particles that may be rapidly catabolized.

 

LOW DENSITY LIPOPROTEINS

 

As noted above the effect of fibrates on LDL-C levels is variable with increases in LDL seen in patients with high TG levels (>400mg/dL) and decreases in LDL-C levels in patients with lower TG levels. In patients with modest elevations in plasma TG levels the clearance of LDL is enhanced (182). The mechanism for this enhanced clearance could be due to a decrease in Apo C-III, as increased levels of this protein inhibits LDL receptor activity (185,186). Additionally, the shift from small dense LDL to large buoyant LDL would enhance the uptake of LDL by the LDL receptor (187). In patients with TG levels > 400mg/dL fibrate therapy decreases LDL clearance (182). Prior to treatment, patients with marked hypertriglyceridemia have hypercatabolism of LDL, which is likely due to increased uptake by the reticuloendothelial system (182). This increased clearance is LDL receptor independent. Treatment with fibrates lowers the plasma TGs leading to normalization of reticuloendothelial cell function and a decrease in LDL clearance resulting in an increase in LDL-C levels with fibrate therapy (182). In addition, the metabolism of VLDL to LDL may be enhanced by fibrates when the TG levels are markedly elevated.

 

Effect of Monotherapy with Fibrates on Cardiovascular Outcomes

 

There have been a number of studies that have examined the effect of monotherapy with a variety of different fibrates on cardiovascular disease. We will describe the major studies below.

 

  • Coronary Drug Project (CDP): CDP conducted between 1966 and 1975, was a randomized, double-blind clinical trial that determined the effect of clofibrate (1.8g/day), dextrothyroxine (6mg/day), two doses of oral estrogen (2.5 or 5mg per day), or immediate release niacin (3 grams/day) vs. placebo in 8,341 men aged 30 to 64 years of age with an electrocardiogram documented myocardial infarction on cardiovascular events and mortality (43). The mean baseline total cholesterol level was 251mg/dL and TG level was 183mg/dL. The two estrogen regimens and dextrothyroxine treatment groups were discontinued early because of increased adverse effects. Clofibrate treatment (n= 1,051) compared to placebo (n= 2,680) also did not demonstrate clinical benefit. The five-year mortality in subjects treated with clofibrate was 20.0% as compared with 20.9% in subjects on placebo therapy (P = 0.55). The results with niacin are discussed above in the section on niacin and cardiovascular outcomes.

 

  • WHO: WHO was a double-blind trial in middle-aged men, age 30-59 years of age, without evidence of heart or other major disease, who were treated with 1.6 grams/day clofibrate (n=5,000) or placebo (n=5,000) for an average of 5.3 years (188). Average serum cholesterol levels were approximately 248mg/dL and a mean reduction of approximately 9 per cent occurred in the clofibrate group. The incidence of ischemic heart disease was decreased by 20% in the clofibrate group compared to the control group (P <0.05). This decrease was confined to non-fatal myocardial infarcts which were reduced by 25% while the incidence of fatal heart attacks and angina was similar in the clofibrate and placebo groups. Importantly, the numbers of deaths and crude mortality rates from all causes were increased in the clofibrate-treated group compared to the control group (P < 0.05). The excess deaths were partially accounted for by increased deaths due to liver, biliary tract, and intestinal disease. There was also an increase in cholecystectomies in subjects treated with clofibrate. Because of increased toxicity clofibrate is no longer available.

 

  • Helsinki Heart Study (HHS): HSS was a randomized double-blind trial in middle aged men age 40-55 years of age without cardiovascular who had non-HDL-C levels greater than or equal to 200mg/dL (152). Subjects were randomized to receive 600mg gemfibrozil twice a day (n=2,051) or placebo (n=2,030) for five years. At initiation of the study total cholesterol was 289mg/dL, HDL-C 47mg/dL, non-HDL-C 242mg/dL, and TGs 176mg/dL. Gemfibrozil caused an increase in HDL-C (approximately 10%) and reductions in total (~10%), LDL-C (~11%), non-HDL-C (~14%), and TG levels (~35%). There were minimal changes in serum lipid levels in the placebo group. Fatal and non-fatal myocardial infarctions and cardiac death were the principal end points and the cumulative rate of these cardiac end points were reduced 34% in the gemfibrozil group (27.3 per 1,000 in the gemfibrozil group vs. 41.4 per 1,000 in the placebo group; P< 0.02). The decrease in cardiovascular disease in the gemfibrozil group became evident in the second year and continued throughout the remainder of the study. There was no difference in mortality between the gemfibrozil and placebo groups. The benefit of gemfibrozil therapy was greatest in participants with elevated TGs and decreased HDL-C levels (189,190). Risk reduction with gemfibrozil was 78% (P = .002) among those with BMI > 26 kg/m2 and dyslipidemia (TGs > ~200mg/dL and HDL-C < 42mg/dL) suggesting that certain types of patients are likely to derive greater benefit from fibrate treatment (191).

 

  • Veterans Affairs High-Density Lipoprotein Cholesterol Intervention Trial (VA-HIT): VA-HIT was a double-blind trial in men with coronary heart disease who had an HDL-C level <40mg/dL and LDL-C level <140mg/dL (153). Subjects were randomized to gemfibrozil 1200mg per day (n=1,264) or placebo (n=1,267) for 5.1 years. Mean lipid levels at study initiation were HDL-C 32mg/dL, LDL-C 111mg/dL, total cholesterol 175mg/dL, and TGs 160mg/dL. At one year, the mean HDL-C level was 6 percent higher, the mean TG level was 31 percent lower, and the mean total cholesterol level was 4 percent lower in the gemfibrozil group than in the placebo group. LDL-C levels did not differ significantly between the groups. The primary study outcome was nonfatal myocardial infarction or death from coronary causes. The primary outcome occurred in 21.7% of patients in the placebo group and 17.3% of patients in the gemfibrozil group (22 percent decrease; P=0.006). A 24% reduction in the combined outcome of death from coronary heart disease, nonfatal myocardial infarction, and stroke was observed in the gemfibrozil group (P< 0.001). There were no significant differences in the rates of coronary revascularization, hospitalization for unstable angina, death from any cause, and cancer. Similar to HHS the beneficial effect of gemfibrozil did not become apparent until approximately two years after treatment. A low HDL-C (<33.5mg/dL) and high TGs (>180mg/dL) at baseline predicted a beneficial response to gemfibrozil therapy (192).

 

  • Bezafibrate Infarction Prevention Study (BIP): BIP was a double-blind study in male and female patients aged 45-74 with a previous myocardial infarction or stable angina (154). Patients were randomized to receive either 400 mg of bezafibrate per day (n=1,548) or a placebo (n=1,542) and were followed for 6.2 years. At the initiation of the study total cholesterol was 212mg/dL, LDL-C was 148mg/dL, HDL-C was 34.6mg/dL, and TGs were145mg/dL. Bezafibrate increased HDL-C by 18% and reduced TGs by 21%. There was a small 7% decrease in LDL-C. The primary end point was fatal or nonfatal myocardial infarction or sudden death. The primary end point occurred in 13. 6% of the bezafibrate group vs. 15.0% of the placebo (9.4% reduction; P=0.26). Total and non-cardiac mortality rates were similar. In a post hoc analysis in the subgroup with high baseline TGs (> or =200 mg/dL), the reduction in the primary end point in the bezafibrate group was 39.5% (P=0.02). Additionally, bezafibrate reduced cardiovascular events in patients with the metabolic syndrome (193). These results again suggest that patients with high TGs are likely to derive benefit from fibrate therapy.

 

  • Leader Trial: The Leader trial was a double blind placebo controlled randomized trial in men age 35 to 92 with lower extremity arterial disease (194,195). Subjects were randomized to bezafibrate 400mg per day (n=783) or placebo (n=785). At baseline total cholesterol levels were 218mg/dL, LDL-C levels 132mg/dL, HDL-C levels 44mg/dL, and TGs 187mg/dL. Bezafibrate therapy reduced total cholesterol levels by 7.6%, LDL-C by 8.1%, and TGs by 23% and increased HDL-C levels by 8%. The primary endpoint of coronary heart disease and strokes was not reduced by bezafibrate treatment. Neither major coronary events nor strokes were significantly reduced.

 .

  • Fenofibrate Intervention and Event Lowering in Diabetes Trial (FIELD): In the FIELD Trial patients with Type 2 diabetes between the ages of 50 and 75 with or without pre-existing cardiovascular disease not taking statin therapy were randomized to fenofibrate 200 mg daily (n=4,895) or placebo (n=4,900) and followed for approximately 5 years (156). At initiation of the study total cholesterol was 196mg/dL, LDL-C was 120mg/dL, HDL-C was 43mg/dL, and TGs were 152mg/dL. Fenofibrate therapy resulted in an 11% decrease in total cholesterol, a 12% decrease in LDL-C, a 29% decrease in TGs, and a 5% increase in HDL-C levels. The primary outcome was coronary events (coronary heart disease death and non-fatal MI), which were reduced by 11% in the fenofibrate group but this difference did not reach statistical significance (p= 0.16). However, there was a 24% decrease in non-fatal MI in the fenofibrate treated group (p=0.01) and a non-significant increase in coronary heart disease mortality. Total cardiovascular disease events (coronary events plus stroke and coronary or carotid revascularization) were reduced 11% (p=0.035). These beneficial effects of fenofibrate therapy on cardiovascular disease were observed in patients without a previous history of cardiovascular disease. In patients with a previous history of cardiovascular disease no benefits were observed. Additionally, the beneficial effect of fenofibrate therapy was seen only in those subjects less than 65 years of age. The beneficial effects of fenofibrate in this study may have been blunted by the increased use of statins in the placebo group, which reduced the differences in lipid levels between the placebo and fenofibrate groups. If one adjusted for the addition of lipid-lowering therapy, fenofibrate reduced the risk of coronary heart disease events by 19% (p=0.01) and of total cardiovascular disease events by 15% (p=0.004). Additionally, many patients in the Field trial did not have elevations in TGs and decreased HDL-C levels. In a post hoc analysis, patients with high TGs 200mg/dL) and low HDL levels (<40mg for men and <50mg/dL for women) derived greater benefit from fenofibrate therapy (196).

 

  • Summary: While the above monotherapy fibrate studies suggest that fibrates reduce cardiovascular event, particularly in patients with high TG and low HDL levels, the results are not as robust or consistent as the beneficial effects of statins on cardiovascular outcomes (5).

 

Effect of Combination Therapy of Fibrates and Statins on Cardiovascular Outcomes

 

Given the marked benefits of statin therapy it is essential to know if adding fibrates to statin therapy further reduces cardiovascular events. Two large trials described below have addressed this key question.

 

  • ACCORD LIPID Trial: The ACCORD-LIPID Trial was designed to determine if the addition of fenofibrate to aggressive statin therapy would result in a further reduction in cardiovascular disease in patients with Type 2 diabetes (197). In this trial, 5,518 patients on statin therapy were randomized to placebo or fenofibrate therapy. The patients had diabetes for approximately 10 years and either had pre-existing cardiovascular disease or were at high risk for developing cardiovascular disease. During the trial, LDL-C levels were approximately 80mg/dL. There was only a small difference in HDL-C with the fenofibrate groups having a mean HDL-C of 41.2mg/dL while the control group had an HDL-C of 40.5mg/dL. Differences in TG levels were somewhat more impressive with the fenofibrate group having a mean TG level of 122mg/dL while the control group had a TG level of 144mg/dL. First occurrence of nonfatal myocardial infarction, nonfatal stroke, or death from cardiovascular causes was the primary outcome and there was no statistical difference between the fenofibrate treated group and the placebo group. Additionally, there were also no statistically significant differences between the groups with regards to any of the secondary outcome measures of cardiovascular disease. Of note, the addition of fenofibrate to statin therapy did not result in an increase in either muscle or liver side effects. On further analysis there was a suggestion of benefit with fenofibrate therapy in the patients in whom the baseline TG levels were elevated (>204mg/dL) and HDL-C levels decreased (<34mg/dL). While this was a negative study, it must be recognized that most of the patients included in this study did not have the lipid profile that would typically lead to treatment with fibrates.

 

  • PROMINENT Trial: The PROMINENT trial studied the effect of pemafibrate, a new selective PPAR-alpha activator, in reducing cardiovascular outcomes in 10,497 patients (66.9% with previous ASCVD) with diabetes (198). This was a double-blind, randomized, controlled trial, in patients with Type 2 diabetes, with mild-to-moderate hypertriglyceridemia (TG level, 200 to 499 mg/dL), LDL-C < 100mg/dL, and HDL-C levels < 40 mg/dL who received either pemafibrate (0.2-mg tablets twice daily) or placebo in addition to statin therapy (96% on statins). The primary end point was a composite of nonfatal MI, ischemic stroke, coronary revascularization, or death from cardiovascular causes. Baseline fasting TG was 271 mg/dL, HDL-C 33 mg/dL, and LDL-C 78 mg/dL. Compared with placebo, pemafibrate decreased TG by 26.2%, while HDL-C increased 5.1% and LDL-C increased 12.3%. Notably non-HDL-C levels were unchanged and Apo B levels increased 4.8%. The primary endpoint was similar in the pemafibrate and placebo group (HR 1.03; 95% CI 0.91 to 1.15). The increase in LDL-C and Apo B levels may have accounted for the failure to reduce cardiovascular events.

 

  • Summary: The results of the ACCORD and PROMINENT trials were disappointing and have greatly reduced the enthusiasm for adding fibrates to statin therapy to cardiovascular events.

 

Effect of Fibrates on Non-Cardiovascular Outcomes

 

DIABETIC RETINOPATHY

 

Small studies in the 1960’s presented suggestive evidence that treatment with clofibrate improved diabetic retinopathy (199,200). Randomized trials have confirmed these observations.

 

The Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) study was a randomized trial in patients with Type 2 diabetes mellitus. Patients were randomly assigned to receive either fenofibrate 200 mg/day (n=4,895) or placebo (n=4,900). Laser treatment for retinopathy was significantly lower in the fenofibrate group than in the placebo group (3.4% patients on fenofibrate vs 4.9% on placebo; p=0.0002) (201). Fenofibrate therapy reduced the need for laser therapy to a similar extent for maculopathy (31% decrease) and for proliferative retinopathy (30% decrease). In the ophthalmology sub-study (n=1,012), the primary endpoint of 2-step progression of retinopathy grade did not differ significantly between the fenofibrate and control groups (9.6% patients on fenofibrate vs 12.3% on placebo; p=0.19). In patients without pre-existing retinopathy there was no difference in progression (11.4% vs 11.7%; p=0.87). However, in patients with pre-existing retinopathy, significantly fewer patients on fenofibrate had a 2-step progression than did those on placebo (3.1% patients vs 14.6%; p=0.004). A composite endpoint of 2-step progression of retinopathy grade, macular edema, or laser treatments was significantly reduced in the fenofibrate group (HR 0.66, 95% CI 0.47-0.94; p=0.022).

 

In the ACCORD Lipids Study a subgroup of participants were evaluated for the progression of diabetic retinopathy by 3 or more steps on the Early Treatment Diabetic Retinopathy Study Severity Scale or the development of diabetic retinopathy necessitating laser photocoagulation or vitrectomy over a four-year period (202). At 4 years, the rates of progression of diabetic retinopathy were 6.5% with fenofibrate therapy (n=806) vs. 10.2% with placebo (n=787) (adjusted odds ratio, 0.60; 95% CI, 0.42 to 0.87; P = 0.006). Of note, this reduction in the progression of diabetic retinopathy was of a similar magnitude as intensive glycemic treatment vs. standard therapy.

 

A double-blind, randomized, placebo-controlled study in 296 patients with type 2 diabetes mellitus evaluated the effect of placebo or etofibrate on diabetic retinopathy (203). After 12 months an improvement in ocular pathology was more frequent in the etofibrate group vs the placebo group ((46% versus 32%; p< 0.001).

 

The MacuFen study was a small double-blind, randomized, placebo-controlled study in 110 subjects with diabetic macular edema who did not require immediate photocoagulation or intraocular treatment (204). Patients were randomized to fenofibric acid or placebo for 1 year. Patients treated with fenofibric acid had a modest improvement in total macular volume that was not statistically significant compared to the placebo group.

 

Taken together these results indicate that fibrates have beneficial effects on the progression of diabetic retinopathy (205). The mechanisms by which fibrates decrease diabetic retinopathy are unknown, and whether decreases in serum TG levels plays an important role is uncertain. Fibrates activate PPAR alpha, which is expressed in the retina (206). Diabetic PPARα KO mice developed more severe DR while overexpression of PPARα in the retina of diabetic rats significantly alleviated diabetes-induced retinal vascular leakage and retinal inflammation, suggesting that fibrates could have direct effects on the retina to reduce diabetic retinopathy (206).

 

DIABETIC KIDNEY DISEASE

 

The Diabetes Atherosclerosis Intervention Study (DAIS) evaluated the effect of fenofibrate therapy (n= 155) vs. placebo (n=159) on changes in urinary albumin excretion in patients with Type 2 diabetes (207). Fenofibrate significantly reduced the worsening of albumin excretion (fenofibrate 8% vs. placebo 18%; P < 0.05). This effect was primarily due to reduced progression from normal albumin excretion to microalbuminuria (fenofibrate 3% vs. 18% placebo; P < 0.001).

 

 In the Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) trial, Type 2 diabetic patients aged 50 to 75 years were randomly assigned to fenofibrate (n = 4,895) or placebo (n = 4,900) for 5 years (208). Fenofibrate reduced urine albumin/creatinine ratio by 24% vs 11% in placebo group (p < 0.001), with 14% less progression and 18% more albuminuria regression (p < 0.001) in the fenofibrate group than in participants on placebo. As expected, fenofibrate therapy acutely increased plasma creatinine levels and decreased eGFR (209). However, over the long-term, the increase in plasma creatinine was lower in the fenofibrate group compared to the placebo group (14% decrease; p=0.01). Similarly, there was a slower annual decrease in eGFR in the fenofibrate group (1.19 vs 2.03 ml/min/1.73 m2annually, p < 0.001). End-stage renal disease, dialysis, renal transplant, and renal death were similar in the fenofibrate and placebo groups, likely due to the small number of events.

 

In the ACCORD-LIPID trial, 5,518 patients on statin therapy were randomized to placebo or fenofibrate therapy (197). The patients had diabetes for approximately 10 years and either had pre-existing cardiovascular disease or were at high risk for developing cardiovascular disease. The post-randomization incidence of microalbuminuria was 38.2% in the fenofibrate group and 41.6% in the placebo group (p=0.01) and post-randomization incidence of macroalbumuria was 10.5% in the fibrate group and 12.3% in the placebo group (p=0.04) indicating a modest reduction in the development of proteinuria in patients treated with fenofibrate (197). There was no significant difference in the incidence of end-stage renal disease or need for dialysis between the fenofibrate group and the placebo group, likely due to the small number of events.

 

A small randomized study in patients with Type 2 diabetes and hypertriglyceridemia compared the effect of fenofibrate (200mg/day) (n=28) vs. no treatment (n=28) on urinary albumin excretion (210). After 180 days urinary albumin/creatine ratio was decreased in the fenofibrate group vs. controls (control -8.15 vs fenofibrate -44.05 mg/g; P<0.05).

 

These studies suggest that fibrates may have a beneficial effect on diabetic kidney disease (211). One should recognize that reducing proteinuria is a surrogate marker and may not indicate a reduction in the development of end stage renal disease. The mechanisms accounting for the decrease in proteinuria are unknown.

 

AMPUTATIONS

 

In the Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) study, patients aged 50-75 years with Type 2 diabetes were randomly assigned to receive fenofibrate 200 mg per day (n=4,895) or matching placebo (n=4,900) for 5 years' duration (212). The risk of first amputation was decreased by 36% (p=0.02) and minor amputation events without known large-vessel disease by 47% (p=0.027) in the fenofibrate treated group (212). The reduction in amputations was independent of glucose control or dyslipidemia. No difference between the risks of major amputations was seen in the placebo and fenofibrate groups. The basis for this reduction in amputations is unknown.

 

GOUT

 

In the Field trial treatment, fenofibrate reduced uric acid levels by 20% and reduced episodes of gout by approximately 50% compared to placebo (HR 0·48, 95% CI 0·37-0·60; p<0·0001) (213). Interestingly, a meta-analysis of fibrate trials found that fenofibrate but not bezafibrate reduced serum uric acid levels suggesting that the reduction in uric acid levels is not a class effect (214).

 

SUMMARY

 

The above studies provide substantial evidence that fibrates have a favorable effect on diabetic microvascular disease (155). While fibrates are not approved specifically for the prevention or treatment of diabetic microvascular disease one should consider these potential beneficial effects when deciding on treatment choices. For example, in a patient with diabetes and microvascular disease and hypertriglyceridemia needing therapy one might elect to use fibrates to lower plasma TGs given their potential beneficial effects on slowing the progression of microvascular disease. 

 

Side Effects

 

RENAL

 

Fibrate therapy leads to an increase in serum creatinine and cystatin C levels (215-217). For example, in the Field Trial serum creatinine levels increased from 0.88mg/dL to 0.99mg/dL, a 12% increase (156). This increase in creatinine has been seen with all fibrates but appears to be less profound with gemfibrozil (215). The increase in cystatin C occurs with fenofibrate but not with other fibrates (216). It must be recognized that this increase in creatinine is reversible on stopping fibrate therapy and does not reflect kidney damage (215). In fact, careful measurements of renal function have not demonstrated a decrease in glomerular filtration rate despite the increase in serum creatinine (209,218,219). As discussed above, studies of renal function in patients with diabetes actually suggests that treatment with fibrates may be protective. The precise mechanism by which fibrates increase serum creatinine levels is unknown.

 

In patients with chronic renal disease fibrates should be used with caution and at lower doses (215). Fibrates are all excreted by the kidneys and thus the excretion of fibrates is decreased in patients with renal dysfunction (215). Therefore, one needs to adjust the fibrate dose depending upon renal function. The National Kidney Foundation recommends the dose adjustments shown in Table 18 (220).

 

Table 18. Fibrate Dose Adjustments in Renal Disease

 

No Kidney Disease

GFR 30-60

GFR < 30

Kidney Transplant

Bezafibrate

400-600mg

200mg

Avoid

Avoid

Ciprofibrate

1000-2000mg

?

Avoid

Avoid

Fenofibrate

150-200mg

40-60mg

Avoid

Avoid

Gemfibrozil

1200mg

1200mg

600mg

600mg

 

GALLBLADDER DISEASE

 

It is clear that clofibrate increases the risk of gallbladder disease. In both the WHO trial and the Coronary Drug Project, cholecystectomies occurred two to three times more often in the patients treated with clofibrate compared to placebo (43,188,221). Whether gemfibrozil, fenofibrate, or other fibrates increases the risk of gallbladder disease is uncertain. In the large randomized outcome studies presented earlier (Effect of fibrates on cardiovascular outcomes section) a statistically significant increase in either gallbladder disease or cholecystectomies were not observed. However, in a sub-study of 450 Helsinki Heart Study participants a trend toward a greater prevalence of gallstones during the study in the gemfibrozil group was observed (7.5% versus 4.9% for the placebo group, a 55% excess for the gemfibrozil group) (Lopid Package Insert). A trend toward a greater incidence of gallbladder surgery was also observed in the gemfibrozil group (17 versus 11 subjects, a 54% excess) (Lopid Package Insert). In a single epidemiological trial fibrate treatment independently correlated with the presence of gallstones with a relative risk of 1.7 (p=0.04) (222).

 

All fibrates alter the composition of bile resulting in an increase in the concentration of cholesterol, which will predispose to the formation of cholesterol gallstones (215). In a comparison of clofibrate and gemfibrozil it was observed that clofibrate resulted in changes in bile composition that would be more lithogenic than gemfibrozil (223).

 

The effect of combining fibrates with statins on the risk of gallbladder disease is unknown.  An increased risk of gallbladder disease or cholecystectomies was not reported in the ACCORD-LIPID trial where fenofibrate was added to statin therapy or the PROMINENT trial where pemafibrate was added to statin therapy (197,198).

 

While it is clear that clofibrate increases the risk of gallbladder disease the effect of other fibrates either as monotherapy or in combination with other drugs is less well defined.

 

PANCREATITIS  

 

In a meta-analysis of 7 fibrate trials involving 40,162 participants conducted over 5.3 years, 144 participants developed pancreatitis (84 assigned to fibrate therapy, 60 assigned to placebo) (RR, 1.39 (95% CI, 1.00-1.95; P = .053) (224). These observations raise the possibility that fibrates may increase the risk of pancreatitis.

 

CANCER

 

A large meta-analysis of 17 randomized controlled trials, involving 44,929 participants, with an average follow-up of 5.2 years has examined if fibrates lead to an increased risk of cancer. No increase in either cancer incidence (RR = 1.02, 95% CI 0.92-1.12) or cancer death (RR = 1.06, 95% CI: 0.92-1.22) was noted with fibrate treatment (225).

 

LIVER DISEASE

 

Fenofibrate has rarely been associated with idiosyncratic hepatotoxicity manifesting as hepatocellular to cholestatic disorders (226). The hepatitis may be acute self-limited or persistent chronic hepatitis. Liver abnormalities are very rare and in large trials such as the FIELD trial described above liver function test abnormalities were similar in the fenofibrate and placebo groups (156).   

 

GLYCEMIC PARAMETERS

 

A meta-analysis of 22 randomized placebo-controlled trials involving a total of 11,402 subjects demonstrated that fibrate therapy significantly decreased fasting plasma glucose, insulin levels, and insulin resistance measured by HOMA-IR, but did not effect HbA1c levels (227).

 

MUSCLE DISORDERS

 

Fibrate monotherapy has been reported to cause myopathy (215). In a large epidemiological study the incidence of hospitalization for rhabdomyolysis per 10,000 person-years for monotherapy with a fibrate was 2.82 (95% CI, 0.58-8.24) while in patients not exposed to lipid lowering drugs the incidence was 0 (95% CI, 0-0.48) (228). The risk of rhabdomyolysis was greater with gemfibrozil therapy than with fenofibrate. Interestingly the incidence of rhabdomyolysis was greater for patients treated with fibrate monotherapy than for patients treated with statin monotherapy (incidence for atorvastatin, pravastatin, or simvastatin was only 0.44 per 10,000 person-years). In an epidemiological study focusing on myopathy similar results were observed (229). The relative risks of myopathy in current users of fibrates and statins compared with nonusers were 42.4 (95% CI = 11.6-170.5) and 7.6 (95% CI = 1.4-41.3), respectively. It should be recognized though that in large randomized clinical trials the risk of muscle symptoms was low in patients treated with fibrates and not dissimilar to that seen in the patients treated with placebo (215). For example, in the Helsinki Heart Study over 2,000 patients were treated and in the VA-HIT over 1,000 patients were treated with gemfibrozil for five years and no cases of  myopathy were reported in either trial (152,153). In the Bezafibrate Infarction Prevention Study, seven patients in the placebo group and five patients in the bezafibrate group reported muscle pain, while CPK levels greater than 2x the upper range of normal was seen in four patients in the bezafibrate group and one patient in the placebo group (154). Finally, in the Field Trial, patients with diabetes were treated with fenofibrate (n=4,895) or placebo (n=4,900) (156). Myositis was observed in one patient treated with placebo and two patients treated with fenofibrate while rhabdomyolysis was observed in one patient treated with placebo and three patients treated with fenofibrate. Elevations in CPK levels values > 10x the upper range of normal were seen in three patients on placebo and 4 patients treated with fenofibrate. Thus, while fibrates can lead to significant muscle dysfunction this is a rare event and appears to occur only slightly more often in patients treated with a fibrate than in patients treated with a placebo. The risk of serious muscle disease appears to be increased in patients with renal failure, hypothyroidism, and in the elderly (215). The mechanism by which fibrates predispose to muscle disorders is unknown.

 

The effect of fibrates in combination with statins on muscle disorders will be discussed in detail in the section on drug interactions below.

 

Drug Interactions

 

STATINS

 

The combination a fibrate and a statin may increase the risk of developing muscle symptoms (215). The degree of risk is dependent on both the specific statin and the specific fibrate that is being used in combination (215). For example, the average incidence per 10,000 person-years for hospitalization for rhabdomyolysis with monotherapy with atorvastatin, pravastatin, or simvastatin was 0.44 (95 % CI, 0.20-0.84); with fibrate alone was 2.82 (95% CI, 0.58-8.24); and with combined therapy of atorvastatin, pravastatin, or simvastatin with a fibrate was 5.98 (95% CI, 0.72-216.0) (228). Of note, the average incidence per 10,000 person-years for hospitalization for rhabdomyolysis with the combination of cerivastatin with a fibrate was 1035 (95% CI, 389-2117), clearly demonstrating an increased risk of the cerivastatin-fibrate combination compared to other statin-fibrate combinations (228). A study by Alsheikh-Ali and colleagues looking at cases of rhabdomyolysis reported to the FDA relative to the total number of prescriptions reached the conclusion that the combination of cerivastatin with a fibrate markedly increased the risk of this complication (230). Additionally, it was noted that the risk of rhabdomyolysis was greater with gemfibrozil compared to fenofibrate and that the combination of cerivastatin and gemfibrozil was particularly toxic (230). Other studies have also noted a marked risk with the combination of cerivastatin and gemfibrozil (231). Cerivastatin is no longer available.

 

Studies comparing the risk of rhabdomyolysis with gemfibrozil-statin combination therapy compared to fenofibrate-statin combination therapy have shown an increased risk with gemfibrozil (215). For example, the number of cases of rhabdomyolysis reported with fenofibrate and statins other than cerivastatin was 0.58 per million prescriptions whereas with gemfibrozil and statins other than cerivastatin was 8.6 per million prescriptions (232). Reviews of the FDA’s adverse events reporting system database have estimated that the risk of myopathy for the combination of gemfibrozil with a statin was much greater than the risk with the combination of fenofibrate with a statin (230,232).  Additionally, studies that employed the combination of gemfibrozil and statins have reported a significant occurrence of muscle related symptoms whereas studies of fenofibrate in combination with statins have not shown an increase in muscle related symptoms (215). For example, the rate of myopathy in over 4,000 patients taking lovastatin was only 0.4% but in patients on the combination of lovastatin and gemfibrozil the frequency increased to 5% (233). In contrast, in the ACCORD-LIPID Trial over 5,000 patients on statin therapy were randomized to fenofibrate or placebo for 4.7 years and no increase in the incidence of muscle related symptoms was observed with fenofibrate therapy (197). Similarly, in the Field Trial approximately 1,000 patients were taking fenofibrate and a statin and with 5 years of follow-up no cases of rhabdomyolysis were reported (156). Finally, a meta-analysis by Geng and colleagues identified 13 randomized trials with 7,712 patients receiving combination fenofibrate-statin therapy compared with statin therapy alone (168). The incidence of elevated creatine kinase levels, muscle-associated adverse events, or withdrawals attributed to muscle dysfunction did not differ significantly between the fenofibrate + statin patients vs. the statin alone patients (168). The American College of Cardiology and American Heart Association Guidelines recommend against using the combination of a statin and gemfibrozil but recognize that the use of a statin and fenofibrate is appropriate under certain circumstances (234).

 

The increased risk of combining gemfibrozil with statins is due to alterations in statin metabolism leading to increases in the serum levels of statins and hence an increased risk of myopathy (215,235). In contrast, fenofibrate does not alter statin metabolism and therefore can be safely combined with statins (Table 19) (235).   

 

Table 19. Effect of Fibrates on Statin Pharmacokinetics (215,235,236)

Statin

Gemfibrozil

Fenofibrate

Atorvastatin

Increase in C-Max by 1.5-Fold

No Change

Simvastatin

Increase in C-Max by 2-Fold

No Change

Pravastatin

Increase in C-Max by 2-Fold

No Change

Rosuvastatin

Increase in C-Max by 2-Fold

No Change

Lovastatin

Increase in C-Max by 2.8-Fold

No Change

Pitavastatin

Increase in C-Max by 41%

Unknown

Fluvastatin

No Change

No Change

  

The explanation for the difference between gemfibrozil and fenofibrate is that gemfibrozil uses the same family of glucuronidation enzymes as the statins thereby inhibiting statin metabolism (215,237). In contrast, fenofibrate uses a different family of glucuronidation enzymes and does not inhibit statin metabolism (215).

 

COUMADIN ANTI-COAGULANTS

 

Gemfibrozil and fenofibrate can potentiate the effect of coumadin anti-coagulants leading to a prolongation of prothrombin time and an increased risk of bleeding. When starting a fibrate in patients on coumadin therapy the dose of coumadin should be decreased and prothrombin times should be closely monitored (Lopid and Tricor Package Inserts).    

 

REPAGLINIDE

 

Gemfibrozil in combination with rapaglinide increases blood levels of rapaglinide and therefore this combination should not be used because of the increased risk of hypoglycemia (Lopid Package Insert).

 

Contraindications

 

Fibrates are contraindicated in patients with severe hepatic dysfunction. Additionally, patients with pre-existing gallstones should not be treated with fibrates. Fenofibrate and gemfibrozil are pregnancy category C drugs and should only be used if the potential benefit justifies the potential risk to the fetus. The combination of gemfibrozil and a statin should be avoided.

 

Conclusions

 

Fibrates are effective drugs in reducing TG levels and modestly increase HDL-C levels. Additionally, they also reduce LDL-C and non-HDL-C levels. Fibrates have a number of side effects and one should avoid using gemfibrozil in combination with statins. In contrast, fenofibrate can be used in combination with statins. Studies have not consistently demonstrated that fibrate monotherapy therapy reduces cardiovascular events and the combination of fibrates and statins in two studies has not been shown to be beneficial. Therefore enthusiasm to use fibrates to reduce cardiovascular events has markedly diminished. In patients with diabetes fibrates appear to slow the progression of microvascular disease. Finally, fibrates are effective in lowering TGs in patients with marked hypertriglyceridemia and while not proven will likely reduce the risk of the development of pancreatitis.

 

VOLANESORSEN

 

Introduction 

 

Volanesorsen (Waylivra) is an antisense oligonucleotide inhibitor of apolipoprotein C-III (apo C-III) mRNA that is approved in Europe for the treatment of familial chylomicronemia syndrome (FCS). This drug has not been approved by the FDA for use in the United States. FCS is a rare metabolic disorder involving the impaired function of lipoprotein lipase (LPL) due to mutations in LPL, Apo C-II, Apo A-V, lipase maturation factor 1, and glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1 (GPIHBP1) (238,239). For a detailed discussion of the diagnosis and treatment of FCS see the following references (238-240).

 

Effect of Volanesorsen on Lipid and Lipoprotein Levels

 

FAMILIAL CHYLOMICRONEMIA SYNDROME (FCS)

 

A double-blind, randomized 52-week trial (APPROACH study) evaluated the ability of volanesorsen (300 mg subcutaneously once weekly) vs. placebo to decrease TG levels in 66 patients with FCS (baseline TGs 2,209mg/dL) (241). The primary end point was the percentage change in fasting TG levels at 3 months. As expected, there was a marked reduction in Apo C-III levels (84% decrease) in the volanesorsen group and a small increase (6%) in the placebo group. Most importantly patients treated with volanesorsen had a 77% decrease at 3 months in TG levels (mean decrease of 1,712 mg/dL) whereas patients receiving placebo had an 18% increase in TG levels. The decrease in TGs in patients treated with volanesorsen persisted for 24 months (242). Significantly, 77% of the patients in the volanesorsen group vs. only 10% of patients in the placebo group had TG levels of less than 750 mg/dL, a level that would greatly reduce the risk of pancreatitis. In addition, patients who received volanesorsen had decreases in levels of chylomicron TG by 83%, apolipoprotein B-48 by 76%, non–HDL-C by 46%, and VLDL-C by 58% and increases in levels of HDL-C by 46%, apolipoprotein A1 by 14%, LDL-C by 136% (note LDL-C increased from 28 to 61 mg/dL), and total apolipoprotein B by 20%.

 

While the APPROACH study was not powered to examine the effect of volanesorsen on pancreatitis, during the study three patients in the placebo group had four episodes of acute pancreatitis, whereas one patient in the volanesorsen group had one episode. In patients with a history of recurrent pancreatitis events (≥ 2 events in the 5 years prior to study, n = 11), a reduction in pancreatitis attacks was seen in patients treated with volanesorsen compared with placebo (none of the 7 patients in the volanesorsen group and 3 of the 4 patients in the placebo group experienced a pancreatitis attack over the 52-week study period).

 

In a retrospective global web-based survey open to all patients with the FCS who received volanesorsen for ≥3 months, 22 patients responded and reported reductions in steatorrhea, pancreatic pain, and constant worry about an attack of pain/ acute pancreatitis (243). The patients also reported that volanesorsen improved overall management of symptoms and reduced interference of FCS with work/school responsibilities. Decreases in the negative impact of FCS on personal, social, and professional life were also reported.

 

HYPERTRIGLYCERIDEMIA

 

A randomized, double-blind, placebo-controlled, study evaluated volanesorsen in patients with hypertriglyceridemia (244). Patients who were not receiving TG-lowering therapy (n=57) were eligible if they had fasting TG level between 350 mg/dL and 2000 mg/dL and were assigned to volanesorsen 100, 200, or 300 mg or placebo. Patients who were receiving a fibrate (n=28) were eligible if they had a fasting TG level between 225 mg/dL and 2000 mg/dL and were randomly assigned to volanesorsen 200 or 300 mg or placebo. The study drug was administered as a single subcutaneous injection once a week for 13 weeks. Baseline TG levels were 581±291 mg/dL in patients not on fibrates and 376±188 mg/dL in patients on fibrates. In patients not on fibrates volanesorsen 300 mg decreased Apo C-III levels by 79.6% vs. an increase of 4.2% in the placebo group (P<0.001) and decreased TG levels by 70.9% compared with an increase of 20.1% in the placebo group (P<0.001). Additionally, HDL-C levels increased by 45.7% from baseline in the 300 mg group, as compared with an increase of 0.7% in the placebo group (P<0.001). LDL-C levels increased from 79.5±29.9 mg/dL to 127.8±44.9 mg/dL with 300 mg of volanesorsen and was associated with an increase in LDL particle size. However, non-HDL-C and total apo B levels remained relatively unchanged and similar to those in the placebo group. Similar changes in Apo C-III, TGs, HDL-C, non-HDL-C, VLDL-C, and total apo B levels were observed in the patients on fibrates treated with volanesorsen. Of note, LDL-C levels did not increase in the patients on fibrates treated with volanesorsen perhaps due to the lower baseline TG levels. 

 

The COMPASS study randomized 113 patients with fasting TGs ≥500 mg/dL (mean TG 1,261mg/dL) to receive either volanesorsen 300 mg or placebo subcutaneously once weekly for 26 weeks (245). Most of these patients had the multifactorial chylomicronemia syndrome but a small number had FCS. A 71% reduction in TGs from baseline after 3 months was observed in patients treated with volanesorsen vs. a 0.9% reduction in placebo-treated patients (P<0.0001). LDL-C levels increased 96% (64 to 111mg/dL), HDL-C increased 61% (25 to 39mg/dL) and non-HDL-C decreased 27% (232 to 158mg/dL) Notably pancreatitis episodes were reduced with 5 events in 3 patients occurring in the placebo group vs. none with volanesorsen treatment (P=0.036). 

 

DIABETES

 

A randomized, double-blind, placebo-controlled trial of volanesorsen 300 mg weekly or placebo was performed in 15 adult patients with type 2 diabetes (HbA1c >7.5%) and hypertriglyceridemia (TG >200 and <500 mg/dL) (246). Treatment with volanesorsen significantly reduced plasma apo C-III (-88%, P = 0.02) and TG (-69%, P = 0.02) levels and raised HDL-C (42%, P = 0.03) without altering LDL-C levels compared with placebo. These changes were accompanied by a 57% improvement in whole-body insulin sensitivity (P < 0.001) and decreases in HbA1c (-0.44%, P = 0.025) 3 months postdosing. The improvement in insulin sensitivity was strongly related to the decrease in plasma apo C-III and TGs.

 

FAMILIAL PARTIAL LIPODYSTROPY (FPL)

 

Patients with FPL were randomized to volanesorsen 300mg weekly (n=21) or placebo (n=19) (247). Median TG level was 781mg/dL in the placebo group and 749mg/dL in the volanesorsen group. Volanesorsen treatment at 3 months resulted in an 88% decrease in TG levels while in the placebo group TG levels decreased by 22% (net difference of −67%; P=0.0009). Non-HDL-HDL-C levels decreased while LDL-C and HDL-C levels increased.

 

Mechanisms Accounting for the Volanesorsen Induced Lipid Effects

 

Volanesorsen binds to apo C-III mRNA leading to increased degradation and thereby inhibits the hepatic synthesis of apo C-III protein resulting in a reduction in plasma apo C-III levels (248,249). Apo C-III has a number of important effects on the metabolism of TG rich lipoproteins (250). Apo C-III is an inhibitor of LPL and therefore decreasing apo C-III levels will enhance LPL activity. In patients with FCS this will not be important because patients with this disorder have defects in components of the LPL complex that result in the inability to increase LPL activity. However, in patients with increased TG levels not due FCS this would accelerate the clearance of TG rich lipoproteins. Studies have also shown that apo C-III stimulates the production and secretion of VLDL by the liver. This effect is also not likely to be of primary importance in patients with FCS as the very high TG levels are primarily due to chylomicrons and not VLDL. However, in other situations increased hepatic secretion of VLDL may be an important contributor to the hypertriglyceridemia. Whether apo C-III regulates chylomicron secretion by the intestine is unknown. Finally, Apo C-III inhibits the binding of TG rich lipoproteins to hepatic LDL receptors and LDL receptor–related protein 1 decreasing the clearance of TG rich lipoprotein particles. A decrease in apo C-III will accelerate the clearance of TG rich lipoproteins, which likely accounts for the ability of volanesorsen to decrease TG levels in patients with FCS.

 

Drug Administration and Pharmacokinetics

 

The recommended starting dose is 285 mg injected subcutaneously once weekly for 3 months after which the dose should be reduced to 285 mg every 2 weeks. If serum TGs decrease by less than 25% or are not below 2000 mg/dL (22.6 mmol/L) after 3 months on volanesorsen 285 mg weekly, treatment should be discontinued (package insert;https://www.ema.europa.eu/en/documents/product-information/waylivra-epar-product-information_en.pdf).

 

After 6 months of treatment one can consider increasing the dose frequency back to 285 mg weekly if the serum TG response has been inadequate and the platelet counts are in the normal range. Patients should return to 285 mg every 2 weeks if the higher 285 mg once weekly dose does not provide a significant additional TG reduction after 9 months (package insert).

 

Effect on Clinical Outcomes

 

As described above in the description of the effect of volanesorsen on lipid/lipoprotein levels in patients with FCS and marked hypertriglyceridemia there is suggestive evidence that lowering the very high TG levels with volanesorsen treatment will reduce the risk of pancreatitis and improve the quality of life.

 

Volanesorsen treatment reduced hepatic fat assessed by MRI in patients with FCS, severe hypertriglyceridemia, and familial partial lipodystrophy (251). The greater the hepatic fat the greater the decrease induced by volanesorsen.

 

The effect of volanesorsen on cardiovascular disease has not been determined. However, epidemiologic studies have demonstrated that increased Apo C-III levels are associated with an increased risk of cardiovascular events (252-254)and coronary artery calcification (255). Moreover, carriers of rare heterozygous loss-of-function mutations in Apo C-III have reduced TG levels and reduced cardiovascular disease risk (256-258). One can speculate that lowering Apo C-III and TG levels with volanesorsen will have beneficial effects on the development of cardiovascular disease.

 

Side Effects

 

Treatment with volanesorsen is very commonly associated with reductions in platelet count in patients with the FCS and may result in thrombocytopenia (package insert; https://www.ema.europa.eu/en/documents/product-information/waylivra-epar-product-information_en.pdf). Platelet counts below 140 x 109/L were observed in 75% of patients treated with volanesorsen vs. 24% of placebo patients. Reductions to below 100 x 109/L were observed in 47% of patients treated with volanesorsen compared with none of the patients in the placebo group. Bleeding secondary to low platelets may occur. Careful monitoring for thrombocytopenia is important during treatment and recommendations for adjustments to monitoring frequency and dosing are shown in table 20 (package insert). Platelet counts recover following drug discontinuation and administration of glucocorticoids where medically indicated.

 

Table 20.  Volanesorsen Monitoring and Treatment Recommendations

Platelet Count (x109/L)

Dose

Monitoring Frequency

Normal (≥140)

Starting dose: Weekly

After 3 months: Every 2 weeks

Every 2 weeks

100-139

Every 2 weeks

Weekly

75-99

Pause treatment for ≥4 weeks and resume treatment after platelet levels ≥ 100 x 109/L

Weekly

50-74

Pause treatment for ≥4 weeks and resume treatment after platelet levels ≥ 100 x 109/L

Every 2-3 days

Less than 50

Discontinue treatment

Glucocorticoids recommended

Daily

 

Renal toxicity has been observed after administration of volanesorsen. Monitoring for evidence of nephrotoxicity by routine urine dipstick is recommended on a quarterly basis. In the case of a positive assessment, one should measure serum creatinine and collect a 24-hour urine collection to quantify the proteinuria and assess creatinine clearance. Treatment should be discontinued if proteinuria ≥ 500 mg/24 hour is present, or an increase in serum creatinine ≥ 0.3 mg/dL that is >ULN occurs, or the creatinine clearance estimated by the CKD-EPI equation is ≤ 30 mL/min/1.73m2(package insert).

 

Elevations of liver enzymes have been observed after administration of volanesorsen. Serum liver enzymes and bilirubin should be monitored every 3 months. Treatment should be discontinued if there is a single increase in ALT or AST > 8 x ULN, or an increase > 5 x ULN, which persists for ≥ 2 weeks, or lesser increases in ALT or AST that are associated with total bilirubin > 2 x ULN or INR > 1.5 (package insert).

 

As expected, injection site reactions are frequently observed and were reported in 82% of patients (erythema, pain, pruritus, or local swelling) (package insert).

 

Contraindications

 

Treatment should not be initiated in patients with thrombocytopenia (platelet count <140 x 109/L). Safety and efficacy have not been established in patients with severe renal disease or patients with hepatic impairment (package insert). There are no data on the use of volanesorsen in pregnant women and it is preferable to avoid the use of volanesorsen during pregnancy (package insert).

 

Drug Interactions

 

Discontinuation of antiplatelet drugs/NSAIDs/anticoagulants should be considered for

platelet levels < 75 x 109/L. Treatment with these products must be discontinued at platelet levels < 50 x 109/L. No other drug interactions have been described (package insert)

 

Conclusions

 

Volanesorsen is a useful drug in patients with the FCS, particularly in patients who have repeated episodes of acute pancreatitis. Whether volanesorsen will be useful for the treatment of less severe hypertriglyceridemia remains to be determined, particularly given its potential side effects. Drugs similar to volanesorsen (Olezarsen) that do not adversely affect platelets are underdevelopment (259).  

 

ALIPOGENE TIPARVOVEC (GLYBERA)

 

Introduction

 

Alipogene tiparvovec is a gene therapy that was approved in Europe for adult patients with Familial Lipoprotein Lipase deficiency and a history of multiple or severe episodes of pancreatitis who have failed dietary therapy (260). The diagnosis of Familial Lipoprotein Lipase with loss of function mutations must be confirmed by genetic testing but patients need to have detectable levels of lipoprotein lipase protein (to avoid immunological reactions) (260). Alipogene tiparvovec is an adeno-associated virus gene therapy that results in the expression of the naturally occurring S447X variant of the human lipoprotein lipase gene that has increased lipoprotein lipase activity compared to “normal” lipoprotein lipase (260). Approximately 20% of Caucasians express this gene variant and these individuals have lower plasma TG levels and an increase in HDL-C levels (261,262). Because of the lack of long-term efficacy alipogene tiparvovec is no longer clinically available.

 

Effect of Alipogene Tiparvovec on Lipid and Lipoprotein Levels

 

In patients with plasma TG levels > 880mg/d, treatment with alipogene tiparvovec resulted in an approximately 40% decrease in fasting plasma TGs with half of the patients having > 40% decrease in fasting plasma TG levels at 3-12 weeks post treatment (263). By week 16-26, fasting TG levels returned to baseline values but chylomicron levels were reduced (263). While fasting TG levels returned to baseline, postprandial TG levels were reduced by approximately 60% suggesting that there are long term effects that are not reflected by fasting TG levels (264). In fact, in some patients treated with alipogene tiparvovec, lipoprotein lipase expression was demonstrated in muscle biopsies at 26 weeks (263).

 

Mechanisms Accounting for the Alipogene Tiparvovec Induced Lipid Effects

 

Gene therapy with alipogene tiparvovec results in the expression of lipoprotein lipase in muscle, which accelerates the clearance of chylomicrons (260,263). Studies have demonstrated a reduced peak level and a reduced area under the curve for postprandial chylomicrons (264).

 

Drug Administration and Pharmacokinetics

 

Alipogene tiparvovec is administered by multiple intramuscularly injections in the legs given at a single visit (260). The number of injections is > 40 and therefore the injections are given under spinal anesthesia (263). From 3 days before administration until 12 weeks after administration patients may be treated with cyclosporine (3mg/kg/day) and mycophenolate (2g/day) and on the day of administration methylprednisolone 1mg/kg) may be administered IV (260,263).

 

Effect on Clinical Outcomes

 

In patients with Familial Lipoprotein Lipase Deficiency the outcome of interest is pancreatitis. In a retrospective study of 19 patients treated with alipogene tiparvovec an approximate 50% decrease in pancreatitis was observed (265). In addition, patients treated with alipogene tiparvovec have reported benefits including discontinuing lipoprotein apheresis, increased energy, and the ability to liberalize their diet, which is difficult to comply with due to the marked limitation in dietary fat (263,266).

 

Conclusions

 

Alipogene tiparvovec may be a useful treatment for the rare patient with Familial Lipoprotein Lipase deficiency but the lack of long-term efficacy and the difficulty of giving the required injections led to this drug being removed from the market.  Because of the rarity of this disorder the information on patients treated with this drug is limited and randomized trials are impossible.

 

EVINACUMAB (EVKEEZA)

 

Introduction

 

Evinacumab is a human monoclonal antibody against angiopoietin-like protein 3 (ANGPTL3). It is approved for the treatment of Homozygous Familial Hypercholesterolemia. Evinacumab decreases LDL-C levels by mechanisms independent of LDL receptor activity. The recommended dose of evinacumab is 15 mg/kg administered by intravenous infusion over 60 minutes every 4 weeks. While it is not approved for TG lowering it is effective in lowering TG levels.

 

Effect on Evinacumab on TG Levels

 

For information on the effect of evinacumab on LDL-C levels see the Endotext chapter on “Cholesterol Lowering Drugs (5). Because of the difficulty in treating severe hypertriglyceridemia, I have focused on evinacumab in this group of patients. Phase 1 studies have shown that various doses of evinacumab lower TG levels in individuals with TG levels between 150-450mg/dL with maximal effects of approximately 80% reductions (267). As one would expect LDL-C and HDL-C levels also decreased in these individuals with modest hypertriglyceridemia.

 

A phase 2 study evaluated evinacumab in three groups of patients with severe hypertriglyceridemia; FCS patients with bi-allelic loss-of-function mutations in the lipoprotein lipase (LPL) pathway (n = 17), multifactorial chylomicronemia syndrome (MFCS) with heterozygous loss-of-function LPL pathway mutations (n = 15), and MFCS without LPL pathway mutations (n = 19) (268). Patients were randomized to evinacumab 15 mg/kg IV or placebo every 4 weeks over 12-weeks. The effect on TG and non-HDL-C levels are shown in table 21. Despite the very small number of patients the results suggest that evinacumab can lower TG levels in patients with MFCS but not in patients with FCS. This result Is not surprising based on the proposed mechanism of action of inhibiting ANGPTL3 (see below).

 

Table 21. Change in Lipid/Lipoprotein Parameters

 

FCS

MFCS/heterozygous LPL pathway mutations

MFCS/ without LPL pathway mutations

 

Placebo (n=5)

Evinacumab (n=12)

Placebo (n=8)

Evinacumab ((n=9)

Placebo (n=5)

Evinacumab (n=14)

Fasting TG

Baseline

3,918mg/dL

3,140mg/dL

1,351mg/dL

1,238mg/dL

1,030mg/dL

1,917mg/dL

% change

−22.9

−27.7

9.4

−64.8*

80.9

−81.7**

Non-HDL-C

Baseline

356mg/dL

345mg/dL

202mg/dL

220mg/dL

209mg/dL

296mg/dL

% change

−15.2

−34.2^

8.0

−31.0^^

48.4

−38.5^^^

*p= 0.0076, **p= 0.0418, ^p= 0.0074, ^^p= 0.0677, ^^^p= 0.1016.

FCS= familial chylomicronemia syndrome, MFCS= multifactorial chylomicronemia syndrome.

 

Mechanism Accounting for the Evinacumab Induced Decrease in TG

 

ANGPTL3 inhibits lipoprotein lipase (LPL) activity thereby slowing the clearance of VLDL and chylomicrons resulting in an increase in plasma triglyceride levels (269,270). Mice deficient in ANGPTL3 have lower plasma triglyceride levels while mice overexpressing ANGPTL3 have elevated plasma triglyceride levels (270). Evinacumab by inhibiting the ability of ANGPTL3 to decrease LPL activity results in an increases in LPL activity, which accelerates the clearance of TG rich lipoproteins decreasing plasma triglyceride levels (270). In patients with FCS who lack a functioning lipoprotein lipase clearance system evinacumab will not accelerate the clearance of TG rich lipoproteins. For information on the mechanism by which evinacumab lowers LDL-C and HDL-C see the Endotext chapter on “Cholesterol Lowering Drugs” (5).

 

Pharmacokinetics and Drug Interactions

 

There are no significant drug interactions.

 

Effect of Evinacumab on Clinical Outcomes

 

There are no cardiovascular outcome studies.

 

Homozygosity for loss-of-function mutations in ANGPTL3 is associated with significantly lower plasma levels of LDL-C, HDL-C, and triglycerides (familial combined hypolipidemia) (270,271). Heterozygous carriers of loss-of-function mutations in ANGPTL3, which occur at a frequency of about 1:300, have significantly lower total cholesterol, LDL-C, and triglyceride levels than noncarriers (270). Moreover, patients carrying loss-of-function variants in ANGPTL3 have a significantly lower risk of coronary artery disease (272,273). Additionally, in an animal model of atherosclerosis treatment with evinacumab decreased atherosclerotic lesion area and necrotic content (272). Taken together these observations suggest that inhibiting ANGPTL3 with evinacumab will reduce cardiovascular disease.

 

Side Effects

 

Serious hypersensitivity reactions have occurred with evinacumab. In clinical trials, 1 (1%) of evinacumab treated patients experienced anaphylaxis vs. 0% of patients who received placebo (package insert).

 

Contraindications

 

Based on animal studies, evinacumab may cause fetal harm when administered to pregnant patients (package insert). Patients should be advised of the potential risks to the fetus of pregnancy. Patients who may become pregnant should be advised to use effective contraception during treatment with evinacumab and for at least 5 months following the last dose.

 

Summary

 

Evinacumab lowers triglyceride levels in patients with severe hypertriglyceridemia due to multifactorial chylomicronemia syndrome and could be useful in selected patients with hypertriglyceridemia. Note it is not approved to treat severe hypertriglyceridemia and administration intravenously every 4 weeks will limit its use to special circumstances.

 

CLINICAL USE OF TRIGLYCERIDE LOWERING DRUGS

 

Marked Hypertriglyceridemia (>500mg/dL); Prevention of Pancreatitis

 

In patients with marked elevations in TG levels (>500-1000mg/dL) the major concern is an increased risk of pancreatitis (274,275). Because of this increased risk it is imperative to lower TG levels. The initial steps are to 1) treat any disease states that could be leading to an elevation in plasma TG levels, 2) if possible, discontinue any drugs that could be leading to an elevation in plasma TGs, and 3) initiate lifestyle changes (Table 22) (2,276).

 

Table 22. Causes of Secondary Hypertriglyceridemia

Lifestyle

Diseases

Medications

Excess calories

Poorly controlled diabetes

Corticosteroids

Excess dietary fat intake

Hypothyroidism

Oral estrogen

Excess simple sugars

Renal disease

Retinoic acid derivatives

Overweight/Obesity

HIV infection

Beta adrenergic blockers

Alcohol intake

Cushing’s syndrome

Thiazide diuretics

Pregnancy

Acromegaly

Protease inhibitors

 

Growth hormone deficiency

Bile acid sequestrants

 

Lipodystrophy

Anti-psychotic drugs

 

Paraproteinemia

Cyclosporine/tacrolimus

 

Nephrotic Syndrome

L-asparaginase

 

Inflammatory Disorders

Interferon alpha 2b

 

 

Cyclophosphamide

 

These initial steps are often sufficient to result in marked reductions in plasma TG levels eliminating the need for TG lowering medications. For example, in patients with diabetes in very poor glycemic control, treatment that results in good glycemic control can markedly lower TG levels (277). Similarly, the restoration of euthyroidism in a hypothyroid patient can also markedly lower lipid levels (278). If these initial steps do not result in a lowering of TGs into an acceptable range, then the use of drugs to lower plasma TG levels is indicated. There have been no randomized controlled trials demonstrating that treatment diminishes pancreatitis but most experienced clinicians believe that lowering TG levels to below 500-1000mg/dL reduces the risk of developing pancreatitis (274,275). The addition of either fibrates or fish oil to lifestyle changes are commonly used to lower markedly elevated TG levels. In some patients, combination therapy is required to lower plasma TGs to an acceptable range. In patients with Familial Chylomicronemia syndrome volanesorsen is a promising therapeutic tool.

 

Moderate Hypertriglyceridemia (150-500mg/dL); Prevention of Cardiovascular Disease

 

In the era of statin therapy, it is uncertain whether lowering TG levels in patients on statin therapy will further reduce cardiovascular events. As discussed in detail in the sections on individual drugs, the studies carried out so far have not shown that adding niacin or fibrates to statin therapy is beneficial with regards to cardiovascular disease. As also discussed, some of the available studies have major limitations because many of the patients in these outcome studies did not have substantial elevations in TGs. Nevertheless, at this time there is little enthusiasm for adding either fibrates or niacin to statins to lower the risk of cardiovascular event.

 

Notably, the REDUCE-IT trial, which tested the effect of high dose EPA (4 grams per day) in patients with elevated TG levels on statin therapy demonstrated a 25% reduction in cardiovascular events. However, the decrease in cardiovascular events was considerably greater than one would expect based on the reduction in TG levels suggesting that the decrease in cardiovascular events was not solely due to lowering TG levels and that other effects of EPA likely played a role. Additionally, as discussed in detail in the section discussing cardiovascular trials in the omega-3-fatty acid section there are concerns that the use of mineral oil as the placebo in the REDUCE-IT trial may have caused harmful effects leading to increased events. Thus, the role of EPA in reducing cardiovascular events is debated with some experts feeling that it is beneficial while others feeling that the evidence for benefit is very weak. Clearly additional studies are required to resolve this controversy. In the meantime, clinicians will need to use their clinical judgement on whether to treat patients with modest elevations in TG levels with EPA (icosapent ethyl; Vascepa) balancing the potential benefits of treatment vs. the potential side effects.

 

Some guidelines use non-HDL-C as a therapeutic goal and thus the use of omega-3-fatty acids and fibrates will often be required to lower TG levels to achieve these non-HDL-C goals. In contrast, other guidelines focus on LDL-C levels and the use of statins and thus de-emphasize the use of omega-3-fatty acids and fibrates. Given the absence of definitive data one needs to use clinical judgement. Consideration should also be given to the use of fenofibrate in hypertriglyceridemic patients with diabetes at high risk for microvascular disease given the studies that have shown that fibrates reduce the microvascular complications of diabetes. Because of the side effects of niacin, the use of niacin to lower TG levels has markedly diminished. In the past we used to use niacin to lower both LDL-C levels and TGs but with the availability of ezetimibe, bempedoic acid, and PCSK9 inhibitors the need to use niacin to lower LDL-C levels has markedly decreased.

 

ACKNOWLEDGEMENTS

 

This work was supported by grants from the Northern California Institute for Research and Education.

 

REFERENCES

 

  1. Geier RR, Tannock LR. Risk of Fasting and Non-Fasting Hypertriglyceridemia in Coronary Vascular Disease and Pancreatitis. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2022.
  2. Feingold KR. 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) 2023.
  3. Feingold KR. Obesity and Dyslipidemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  4. Feingold KR. The Effect of Diet on Cardiovascular Disease and Lipid and Lipoprotein Levels. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  5. 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) 2021.
  6. Altschul R, Hoffer A, Stephen JD. Influence of nicotinic acid on serum cholesterol in man. Arch Biochem Biophys1955; 54:558-559
  7. Cooper DL, Murrell DE, Roane DS, Harirforoosh S. Effects of formulation design on niacin therapeutics: mechanism of action, metabolism, and drug delivery. Int J Pharm 2015; 490:55-64
  8. Song WL, FitzGerald GA. Niacin, an old drug with a new twist. J Lipid Res 2013; 54:2586-2594
  9. Julius U. Niacin as antidyslipidemic drug. Can J Physiol Pharmacol 2015; 93:1043-1054
  10. Morgan JM, Capuzzi DM, Baksh RI, Intenzo C, Carey CM, Reese D, Walker K. Effects of extended-release niacin on lipoprotein subclass distribution. Am J Cardiol 2003; 91:1432-1436
  11. Birjmohun RS, Hutten BA, Kastelein JJ, Stroes ES. Efficacy and safety of high-density lipoprotein cholesterol-increasing compounds: a meta-analysis of randomized controlled trials. J Am Coll Cardiol 2005; 45:185-197
  12. Knopp RH, Alagona P, Davidson M, Goldberg AC, Kafonek SD, Kashyap M, Sprecher D, Superko HR, Jenkins S, Marcovina S. Equivalent efficacy of a time-release form of niacin (Niaspan) given once-a-night versus plain niacin in the management of hyperlipidemia. Metabolism 1998; 47:1097-1104
  13. Sahebkar A, Reiner Z, Simental-Mendia LE, Ferretti G, Cicero AF. Effect of extended-release niacin on plasma lipoprotein(a) levels: A systematic review and meta-analysis of randomized placebo-controlled trials. Metabolism2016; 65:1664-1678
  14. Goldberg AC. A meta-analysis of randomized controlled studies on the effects of extended-release niacin in women. Am J Cardiol 2004; 94:121-124
  15. Ballantyne CM, Davidson MH, McKenney J, Keller LH, Bajorunas DR, Karas RH. Comparison of the safety and efficacy of a combination tablet of niacin extended release and simvastatin vs simvastatin monotherapy in patients with increased non-HDL cholesterol (from the SEACOAST I study). Am J Cardiol 2008; 101:1428-1436
  16. Fazio S, Guyton JR, Polis AB, Adewale AJ, Tomassini JE, Ryan NW, Tershakovec AM. Long-term safety and efficacy of triple combination ezetimibe/simvastatin plus extended-release niacin in patients with hyperlipidemia. Am J Cardiol 2010; 105:487-494
  17. Shearer GC, Pottala JV, Hansen SN, Brandenburg V, Harris WS. Effects of prescription niacin and omega-3 fatty acids on lipids and vascular function in metabolic syndrome: a randomized controlled trial. J Lipid Res 2012; 53:2429-2435
  18. Pradhan B, Neopane A, Karki S, Karki DB. Effectiveness of nicotinic acid and bezafibrate alone and in combination for reducing serum triglyceride level. Kathmandu Univ Med J (KUMJ) 2005; 3:411-414
  19. Balasubramanyam A, Coraza I, Smith EO, Scott LW, Patel P, Iyer D, Taylor AA, Giordano TP, Sekhar RV, Clark P, Cuevas-Sanchez E, Kamble S, Ballantyne CM, Pownall HJ. Combination of niacin and fenofibrate with lifestyle changes improves dyslipidemia and hypoadiponectinemia in HIV patients on antiretroviral therapy: results of "heart positive," a randomized, controlled trial. J Clin Endocrinol Metab 2011; 96:2236-2247
  20. Carlson LA. Studies on the effect of nicotinic acid on catecholamine stimulated lipolysis in adipose tissue in vitro. Acta Med Scand 1963; 173:719-722
  21. Carlson LA, Oro L. The effect of nicotinic acid on the plasma free fatty acid; demonstration of a metabolic type of sympathicolysis. Acta Med Scand 1962; 172:641-645
  22. Kamanna VS, Ganji SH, Kashyap ML. Recent advances in niacin and lipid metabolism. Curr Opin Lipidol 2013; 24:239-245
  23. Tunaru S, Kero J, Schaub A, Wufka C, Blaukat A, Pfeffer K, Offermanns S. PUMA-G and HM74 are receptors for nicotinic acid and mediate its anti-lipolytic effect. Nat Med 2003; 9:352-355
  24. Wise A, Foord SM, Fraser NJ, Barnes AA, Elshourbagy N, Eilert M, Ignar DM, Murdock PR, Steplewski K, Green A, Brown AJ, Dowell SJ, Szekeres PG, Hassall DG, Marshall FH, Wilson S, Pike NB. Molecular identification of high and low affinity receptors for nicotinic acid. J Biol Chem 2003; 278:9869-9874
  25. Wang W, Basinger A, Neese RA, Christiansen M, Hellerstein MK. Effects of nicotinic acid on fatty acid kinetics, fuel selection, and pathways of glucose production in women. Am J Physiol Endocrinol Metab 2000; 279:E50-59
  26. Lauring B, Taggart AK, Tata JR, Dunbar R, Caro L, Cheng K, Chin J, Colletti SL, Cote J, Khalilieh S, Liu J, Luo WL, Maclean AA, Peterson LB, Polis AB, Sirah W, Wu TJ, Liu X, Jin L, Wu K, Boatman PD, Semple G, Behan DP, Connolly DT, Lai E, Wagner JA, Wright SD, Cuffie C, Mitchel YB, Rader DJ, Paolini JF, Waters MG, Plump A. Niacin lipid efficacy is independent of both the niacin receptor GPR109A and free fatty acid suppression. Sci Transl Med 2012; 4:148ra115
  27. Ganji SH, Tavintharan S, Zhu D, Xing Y, Kamanna VS, Kashyap ML. Niacin noncompetitively inhibits DGAT2 but not DGAT1 activity in HepG2 cells. J Lipid Res 2004; 45:1835-1845
  28. Grundy SM, Mok HY, Zech L, Berman M. Influence of nicotinic acid on metabolism of cholesterol and triglycerides in man. J Lipid Res 1981; 22:24-36
  29. Blond E, Rieusset J, Alligier M, Lambert-Porcheron S, Bendridi N, Gabert L, Chetiveaux M, Debard C, Chauvin MA, Normand S, Roth H, de Gouville AC, Krempf M, Vidal H, Goudable J, Laville M, Niacin" Study G. Nicotinic acid effects on insulin sensitivity and hepatic lipid metabolism: an in vivo to in vitro study. Horm Metab Res 2014; 46:390-396
  30. Hernandez C, Molusky M, Li Y, Li S, Lin JD. Regulation of hepatic ApoC3 expression by PGC-1beta mediates hypolipidemic effect of nicotinic acid. Cell Metab 2010; 12:411-419
  31. Zhang LH, Kamanna VS, Zhang MC, Kashyap ML. Niacin inhibits surface expression of ATP synthase beta chain in HepG2 cells: implications for raising HDL. J Lipid Res 2008; 49:1195-1201
  32. Jin FY, Kamanna VS, Kashyap ML. Niacin decreases removal of high-density lipoprotein apolipoprotein A-I but not cholesterol ester by Hep G2 cells. Implication for reverse cholesterol transport. Arterioscler Thromb Vasc Biol1997; 17:2020-2028
  33. Blum CB, Levy RI, Eisenberg S, Hall M, 3rd, Goebel RH, Berman M. High density lipoprotein metabolism in man. J Clin Invest 1977; 60:795-807
  34. Shepherd J, Packard CJ, Patsch JR, Gotto AM, Jr., Taunton OD. Effects of nicotinic acid therapy on plasma high density lipoprotein subfraction distribution and composition and on apolipoprotein A metabolism. J Clin Invest1979; 63:858-867
  35. Lamon-Fava S, Diffenderfer MR, Barrett PH, Buchsbaum A, Nyaku M, Horvath KV, Asztalos BF, Otokozawa S, Ai M, Matthan NR, Lichtenstein AH, Dolnikowski GG, Schaefer EJ. Extended-release niacin alters the metabolism of plasma apolipoprotein (Apo) A-I and ApoB-containing lipoproteins. Arterioscler Thromb Vasc Biol 2008; 28:1672-1678
  36. Rubic T, Trottmann M, Lorenz RL. Stimulation of CD36 and the key effector of reverse cholesterol transport ATP-binding cassette A1 in monocytoid cells by niacin. Biochem Pharmacol 2004; 67:411-419
  37. Zhang LH, Kamanna VS, Ganji SH, Xiong XM, Kashyap ML. Niacin increases HDL biogenesis by enhancing DR4-dependent transcription of ABCA1 and lipidation of apolipoprotein A-I in HepG2 cells. J Lipid Res 2012; 53:941-950
  38. van der Hoorn JW, de Haan W, Berbee JF, Havekes LM, Jukema JW, Rensen PC, Princen HM. Niacin increases HDL by reducing hepatic expression and plasma levels of cholesteryl ester transfer protein in APOE*3Leiden.CETP mice. Arterioscler Thromb Vasc Biol 2008; 28:2016-2022
  39. Seed M, O'Connor B, Perombelon N, O'Donnell M, Reaveley D, Knight BL. The effect of nicotinic acid and acipimox on lipoprotein(a) concentration and turnover. Atherosclerosis 1993; 101:61-68
  40. Croyal M, Ouguerram K, Passard M, Ferchaud-Roucher V, Chetiveaux M, Billon-Crossouard S, de Gouville AC, Lambert G, Krempf M, Nobecourt E. Effects of Extended-Release Nicotinic Acid on Apolipoprotein (a) Kinetics in Hypertriglyceridemic Patients. Arterioscler Thromb Vasc Biol 2015; 35:2042-2047
  41. Chennamsetty I, Kostner KM, Claudel T, Vinod M, Frank S, Weiss TS, Trauner M, Kostner GM. Nicotinic acid inhibits hepatic APOA gene expression: studies in humans and in transgenic mice. J Lipid Res 2012; 53:2405-2412
  42. Pieper JA. Overview of niacin formulations: differences in pharmacokinetics, efficacy, and safety. Am J Health Syst Pharm 2003; 60:S9-14; quiz S25
  43. Clofibrate and niacin in coronary heart disease. JAMA 1975; 231:360-381
  44. Canner PL, Berge KG, Wenger NK, Stamler J, Friedman L, Prineas RJ, Friedewald W. Fifteen year mortality in Coronary Drug Project patients: long-term benefit with niacin. J Am Coll Cardiol 1986; 8:1245-1255
  45. Canner PL, Furberg CD, Terrin ML, McGovern ME. Benefits of niacin by glycemic status in patients with healed myocardial infarction (from the Coronary Drug Project). Am J Cardiol 2005; 95:254-257
  46. Canner PL, Furberg CD, McGovern ME. Benefits of niacin in patients with versus without the metabolic syndrome and healed myocardial infarction (from the Coronary Drug Project). Am J Cardiol 2006; 97:477-479
  47. Carlson LA, Rosenhamer G. Reduction of mortality in the Stockholm Ischaemic Heart Disease Secondary Prevention Study by combined treatment with clofibrate and nicotinic acid. Acta Med Scand 1988; 223:405-418
  48. Carlson LA, Danielson M, Ekberg I, Klintemar B, Rosenhamer G. Reduction of myocardial reinfarction by the combined treatment with clofibrate and nicotinic acid. Atherosclerosis 1977; 28:81-86
  49. AIM-HIGH Investigators, Boden WE, Probstfield JL, Anderson T, Chaitman BR, Desvignes-Nickens P, Koprowicz K, McBride R, Teo K, Weintraub W. Niacin in patients with low HDL cholesterol levels receiving intensive statin therapy. N Engl J Med 2011; 365:2255-2267
  50. Guyton JR, Slee AE, Anderson T, Fleg JL, Goldberg RB, Kashyap ML, Marcovina SM, Nash SD, O'Brien KD, Weintraub WS, Xu P, Zhao XQ, Boden WE. Relationship of lipoproteins to cardiovascular events: the AIM-HIGH Trial (Atherothrombosis Intervention in Metabolic Syndrome With Low HDL/High Triglycerides and Impact on Global Health Outcomes). J Am Coll Cardiol 2013; 62:1580-1584
  51. Hps Thrive Collaborative Group, Landray MJ, Haynes R, Hopewell JC, Parish S, Aung T, Tomson J, Wallendszus K, Craig M, Jiang L, Collins R, Armitage J. Effects of extended-release niacin with laropiprant in high-risk patients. N Engl J Med 2014; 371:203-212
  52. Song WL, Stubbe J, Ricciotti E, Alamuddin N, Ibrahim S, Crichton I, Prempeh M, Lawson JA, Wilensky RL, Rasmussen LM, Pure E, FitzGerald GA. Niacin and biosynthesis of PGD(2)by platelet COX-1 in mice and humans. J Clin Invest 2012; 122:1459-1468
  53. Cashin-Hemphill L, Mack WJ, Pogoda JM, Sanmarco ME, Azen SP, Blankenhorn DH. Beneficial effects of colestipol-niacin on coronary atherosclerosis. A 4-year follow-up. JAMA 1990; 264:3013-3017
  54. Brown G, Albers JJ, Fisher LD, Schaefer SM, Lin JT, Kaplan C, Zhao XQ, Bisson BD, Fitzpatrick VF, Dodge HT. Regression of coronary artery disease as a result of intensive lipid-lowering therapy in men with high levels of apolipoprotein B. N Engl J Med 1990; 323:1289-1298
  55. 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
  56. Brown BG, Zhao XQ, Chait A, Fisher LD, Cheung MC, Morse JS, Dowdy AA, Marino EK, Bolson EL, Alaupovic P, Frohlich J, Albers JJ. Simvastatin and niacin, antioxidant vitamins, or the combination for the prevention of coronary disease. N Engl J Med 2001; 345:1583-1592
  57. Whitney EJ, Krasuski RA, Personius BE, Michalek JE, Maranian AM, Kolasa MW, Monick E, Brown BG, Gotto AM, Jr. A randomized trial of a strategy for increasing high-density lipoprotein cholesterol levels: effects on progression of coronary heart disease and clinical events. Ann Intern Med 2005; 142:95-104
  58. Sacks FM, Pasternak RC, Gibson CM, Rosner B, Stone PH. Effect on coronary atherosclerosis of decrease in plasma cholesterol concentrations in normocholesterolaemic patients. Harvard Atherosclerosis Reversibility Project (HARP) Group. Lancet 1994; 344:1182-1186
  59. Taylor AJ, Lee HJ, Sullenberger LE. The effect of 24 months of combination statin and extended-release niacin on carotid intima-media thickness: ARBITER 3. Curr Med Res Opin 2006; 22:2243-2250
  60. Taylor AJ, Sullenberger LE, Lee HJ, Lee JK, Grace KA. Arterial Biology for the Investigation of the Treatment Effects of Reducing Cholesterol (ARBITER) 2: a double-blind, placebo-controlled study of extended-release niacin on atherosclerosis progression in secondary prevention patients treated with statins. Circulation 2004; 110:3512-3517
  61. Taylor AJ, Villines TC, Stanek EJ, Devine PJ, Griffen L, Miller M, Weissman NJ, Turco M. Extended-release niacin or ezetimibe and carotid intima-media thickness. N Engl J Med 2009; 361:2113-2122
  62. Thoenes M, Oguchi A, Nagamia S, Vaccari CS, Hammoud R, Umpierrez GE, Khan BV. The effects of extended-release niacin on carotid intimal media thickness, endothelial function and inflammatory markers in patients with the metabolic syndrome. Int J Clin Pract 2007; 61:1942-1948
  63. Lee JM, Robson MD, Yu LM, Shirodaria CC, Cunnington C, Kylintireas I, Digby JE, Bannister T, Handa A, Wiesmann F, Durrington PN, Channon KM, Neubauer S, Choudhury RP. Effects of high-dose modified-release nicotinic acid on atherosclerosis and vascular function: a randomized, placebo-controlled, magnetic resonance imaging study. J Am Coll Cardiol 2009; 54:1787-1794
  64. Dunbar RL, Gelfand JM. Seeing red: flushing out instigators of niacin-associated skin toxicity. J Clin Invest 2010; 120:2651-2655
  65. Dunn RT, Ford MA, Rindone JP, Kwiecinski FA. Low-Dose Aspirin and Ibuprofen Reduce the Cutaneous Reactions Following Niacin Administration. Am J Ther 1995; 2:478-480
  66. McKenney JM, Proctor JD, Harris S, Chinchili VM. A comparison of the efficacy and toxic effects of sustained- vs immediate-release niacin in hypercholesterolemic patients. JAMA 1994; 271:672-677
  67. Henkin Y, Johnson KC, Segrest JP. Rechallenge with crystalline niacin after drug-induced hepatitis from sustained-release niacin. JAMA 1990; 264:241-243
  68. Kashyap ML, Ganji S, Nakra NK, Kamanna VS. Niacin for treatment of nonalcoholic fatty liver disease (NAFLD): novel use for an old drug? J Clin Lipidol 2019; 13:873-879
  69. Anderson TJ, Boden WE, Desvigne-Nickens P, Fleg JL, Kashyap ML, McBride R, Probstfield JL, AIM_HIGH Investigators. Safety profile of extended-release niacin in the AIM-HIGH trial. N Engl J Med 2014; 371:288-290
  70. Miettinen TA, Taskinen MR, Pelkonen R, Nikkila EA. Glucose tolerance and plasma insulin in man during acute and chronic administration of nicotinic acid. Acta Med Scand 1969; 186:247-253
  71. Poynten AM, Gan SK, Kriketos AD, O'Sullivan A, Kelly JJ, Ellis BA, Chisholm DJ, Campbell LV. Nicotinic acid-induced insulin resistance is related to increased circulating fatty acids and fat oxidation but not muscle lipid content. Metabolism 2003; 52:699-704
  72. Goldberg RB, Bittner VA, Dunbar RL, Fleg JL, Grunberger G, Guyton JR, Leiter LA, McBride R, Robinson JG, Simmons DL, Wysham C, Xu P, Boden WE. Effects of Extended-Release Niacin Added to Simvastatin/Ezetimibe on Glucose and Insulin Values in AIM-HIGH. Am J Med 2016; 129:753 e713-722
  73. Goldie C, Taylor AJ, Nguyen P, McCoy C, Zhao XQ, Preiss D. Niacin therapy and the risk of new-onset diabetes: a meta-analysis of randomised controlled trials. Heart 2016; 102:198-203
  74. Elam MB, Hunninghake DB, Davis KB, Garg R, Johnson C, Egan D, Kostis JB, Sheps DS, Brinton EA. Effect of niacin on lipid and lipoprotein levels and glycemic control in patients with diabetes and peripheral arterial disease: the ADMIT study: A randomized trial. Arterial Disease Multiple Intervention Trial. JAMA 2000; 284:1263-1270
  75. Garg A, Grundy SM. Nicotinic acid as therapy for dyslipidemia in non-insulin-dependent diabetes mellitus. JAMA1990; 264:723-726
  76. Ding Y, Li Y, Wen A. Effect of niacin on lipids and glucose in patients with type 2 diabetes: A meta-analysis of randomized, controlled clinical trials. Clin Nutr 2015; 34:838-844
  77. Gershon SL, Fox IH. Pharmacologic effects of nicotinic acid on human purine metabolism. J Lab Clin Med 1974; 84:179-186
  78. Gagne JJ, Houstoun M, Reichman ME, Hampp C, Marshall JH, Toh S. Safety assessment of niacin in the US Food and Drug Administration's mini-sentinel system. Pharmacoepidemiol Drug Saf 2018; 27:30-37
  79. Domanico D, Verboschi F, Altimari S, Zompatori L, Vingolo EM. Ocular Effects of Niacin: A Review of the Literature. Med Hypothesis Discov Innov Ophthalmol 2015; 4:64-71
  80. Zargar A, Ito MK. Long chain omega-3 dietary supplements: a review of the National Library of Medicine Herbal Supplement Database. Metab Syndr Relat Disord 2011; 9:255-271
  81. Kleiner AC, Cladis DP, Santerre CR. A comparison of actual versus stated label amounts of EPA and DHA in commercial omega-3 dietary supplements in the United States. J Sci Food Agric 2015; 95:1260-1267
  82. Wendland E, Farmer A, Glasziou P, Neil A. Effect of alpha linolenic acid on cardiovascular risk markers: a systematic review. Heart 2006; 92:166-169
  83. Eslick GD, Howe PR, Smith C, Priest R, Bensoussan A. Benefits of fish oil supplementation in hyperlipidemia: a systematic review and meta-analysis. Int J Cardiol 2009; 136:4-16
  84. Balk EM, Lichtenstein AH, Chung M, Kupelnick B, Chew P, Lau J. Effects of omega-3 fatty acids on serum markers of cardiovascular disease risk: a systematic review. Atherosclerosis 2006; 189:19-30
  85. Hartweg J, Perera R, Montori V, Dinneen S, Neil HA, Farmer A. Omega-3 polyunsaturated fatty acids (PUFA) for type 2 diabetes mellitus. Cochrane Database Syst Rev 2008:CD003205
  86. Zheng T, Zhao J, Wang Y, Liu W, Wang Z, Shang Y, Zhang W, Zhang Y, Zhong M. The limited effect of omega-3 polyunsaturated fatty acids on cardiovascular risk in patients with impaired glucose metabolism: a meta-analysis. Clin Biochem 2014; 47:369-377
  87. Chi H, Lin X, Huang H, Zheng X, Li T, Zou Y. Omega-3 fatty acid supplementation on lipid profiles in dialysis patients: meta-analysis. Arch Med Res 2014; 45:469-477
  88. Pei J, Zhao Y, Huang L, Zhang X, Wu Y. The effect of n-3 polyunsaturated fatty acids on plasma lipids and lipoproteins in patients with chronic renal failure--a meta-analysis of randomized controlled trials. J Ren Nutr 2012; 22:525-532
  89. Zhu W, Dong C, Du H, Zhang H, Chen J, Hu X, Hu F. Effects of fish oil on serum lipid profile in dialysis patients: a systematic review and meta-analysis of randomized controlled trials. Lipids Health Dis 2014; 13:127
  90. Hall AV, Parbtani A, Clark WF, Spanner E, Huff MW, Philbrick DJ, Holub BJ. Omega-3 fatty acid supplementation in primary nephrotic syndrome: effects on plasma lipids and coagulopathy. J Am Soc Nephrol 1992; 3:1321-1329
  91. Spadaro L, Magliocco O, Spampinato D, Piro S, Oliveri C, Alagona C, Papa G, Rabuazzo AM, Purrello F. Effects of n-3 polyunsaturated fatty acids in subjects with nonalcoholic fatty liver disease. Dig Liver Dis 2008; 40:194-199
  92. Oliveira JM, Rondo PH. Omega-3 fatty acids and hypertriglyceridemia in HIV-infected subjects on antiretroviral therapy: systematic review and meta-analysis. HIV Clin Trials 2011; 12:268-274
  93. De Truchis P, Kirstetter M, Perier A, Meunier C, Zucman D, Force G, Doll J, Katlama C, Rozenbaum W, Masson H, Gardette J, Melchior JC. Reduction in triglyceride level with N-3 polyunsaturated fatty acids in HIV-infected patients taking potent antiretroviral therapy: a randomized prospective study. J Acquir Immune Defic Syndr 2007; 44:278-285
  94. Harris WS, Ginsberg HN, Arunakul N, Shachter NS, Windsor SL, Adams M, Berglund L, Osmundsen K. Safety and efficacy of Omacor in severe hypertriglyceridemia. J Cardiovasc Risk 1997; 4:385-391
  95. Maki KC, Orloff DG, Nicholls SJ, Dunbar RL, Roth EM, Curcio D, Johnson J, Kling D, Davidson MH. A highly bioavailable omega-3 free fatty acid formulation improves the cardiovascular risk profile in high-risk, statin-treated patients with residual hypertriglyceridemia (the ESPRIT trial). Clin Ther 2013; 35:1400-1411 e1401-1403
  96. Pownall HJ, Brauchi D, Kilinc C, Osmundsen K, Pao Q, Payton-Ross C, Gotto AM, Jr., Ballantyne CM. Correlation of serum triglyceride and its reduction by omega-3 fatty acids with lipid transfer activity and the neutral lipid compositions of high-density and low-density lipoproteins. Atherosclerosis 1999; 143:285-297
  97. Calabresi L, Donati D, Pazzucconi F, Sirtori CR, Franceschini G. Omacor in familial combined hyperlipidemia: effects on lipids and low density lipoprotein subclasses. Atherosclerosis 2000; 148:387-396
  98. Minihane AM, Khan S, Leigh-Firbank EC, Talmud P, Wright JW, Murphy MC, Griffin BA, Williams CM. ApoE polymorphism and fish oil supplementation in subjects with an atherogenic lipoprotein phenotype. Arterioscler Thromb Vasc Biol 2000; 20:1990-1997
  99. Haglund O, Mehta JL, Saldeen T. Effects of fish oil on some parameters of fibrinolysis and lipoprotein(a) in healthy subjects. Am J Cardiol 1994; 74:189-192
  100. Beil FU, Terres W, Orgass M, Greten H. Dietary fish oil lowers lipoprotein(a) in primary hypertriglyceridemia. Atherosclerosis 1991; 90:95-97
  101. Herrmann W, Biermann J, Kostner GM. Comparison of effects of N-3 to N-6 fatty acids on serum level of lipoprotein(a) in patients with coronary artery disease. Am J Cardiol 1995; 76:459-462
  102. Shinozaki K, Kambayashi J, Kawasaki T, Uemura Y, Sakon M, Shiba E, Shibuya T, Nakamura T, Mori T. The long-term effect of eicosapentaenoic acid on serum levels of lipoprotein (a) and lipids in patients with vascular disease. J Atheroscler Thromb 1996; 2:107-109
  103. Eritsland J, Arnesen H, Berg K, Seljeflot I, Abdelnoor M. Serum Lp(a) lipoprotein levels in patients with coronary artery disease and the influence of long-term n-3 fatty acid supplementation. Scand J Clin Lab Invest 1995; 55:295-300
  104. Davidson MH, Stein EA, Bays HE, Maki KC, Doyle RT, Shalwitz RA, Ballantyne CM, Ginsberg HN. Efficacy and tolerability of adding prescription omega-3 fatty acids 4 g/d to simvastatin 40 mg/d in hypertriglyceridemic patients: an 8-week, randomized, double-blind, placebo-controlled study. Clin Ther 2007; 29:1354-1367
  105. Bays HE, Ballantyne CM, Kastelein JJ, Isaacsohn JL, Braeckman RA, Soni PN. Eicosapentaenoic acid ethyl ester (AMR101) therapy in patients with very high triglyceride levels (from the Multi-center, plAcebo-controlled, Randomized, double-blINd, 12-week study with an open-label Extension [MARINE] trial). Am J Cardiol 2011; 108:682-690
  106. Ballantyne CM, Bays HE, Kastelein JJ, Stein E, Isaacsohn JL, Braeckman RA, Soni PN. Efficacy and safety of eicosapentaenoic acid ethyl ester (AMR101) therapy in statin-treated patients with persistent high triglycerides (from the ANCHOR study). Am J Cardiol 2012; 110:984-992
  107. Kastelein JJ, Maki KC, Susekov A, Ezhov M, Nordestgaard BG, Machielse BN, Kling D, Davidson MH. Omega-3 free fatty acids for the treatment of severe hypertriglyceridemia: the EpanoVa fOr Lowering Very high triglyceridEs (EVOLVE) trial. J Clin Lipidol 2014; 8:94-106
  108. Wei MY, Jacobson TA. Effects of eicosapentaenoic acid versus docosahexaenoic acid on serum lipids: a systematic review and meta-analysis. Curr Atheroscler Rep 2011; 13:474-483
  109. Tatsuno I, Saito Y, Kudou K, Ootake J. Efficacy and safety of TAK-085 compared with eicosapentaenoic acid in Japanese subjects with hypertriglyceridemia undergoing lifestyle modification: the omega-3 fatty acids randomized double-blind (ORD) study. J Clin Lipidol 2013; 7:199-207
  110. Tatsuno I, Saito Y, Kudou K, Ootake J. Long-term safety and efficacy of TAK-085 in Japanese subjects with hypertriglyceridemia undergoing lifestyle modification: the omega-3 fatty acids randomized long-term (ORL) study. J Clin Lipidol 2013; 7:615-625
  111. Roth EM, Bays HE, Forker AD, Maki KC, Carter R, Doyle RT, Stein EA. Prescription omega-3 fatty acid as an adjunct to fenofibrate therapy in hypertriglyceridemic subjects. J Cardiovasc Pharmacol 2009; 54:196-203
  112. Shearer GC, Savinova OV, Harris WS. Fish oil -- how does it reduce plasma triglycerides? Biochim Biophys Acta2012; 1821:843-851
  113. Harris WS, Bulchandani D. Why do omega-3 fatty acids lower serum triglycerides? Curr Opin Lipidol 2006; 17:387-393
  114. Ooi EM, Watts GF, Ng TW, Barrett PH. Effect of dietary Fatty acids on human lipoprotein metabolism: a comprehensive update. Nutrients 2015; 7:4416-4425
  115. Davidson MH. Mechanisms for the hypotriglyceridemic effect of marine omega-3 fatty acids. Am J Cardiol 2006; 98:27i-33i
  116. Wang H, Chen X, Fisher EA. N-3 fatty acids stimulate intracellular degradation of apoprotein B in rat hepatocytes. J Clin Invest 1993; 91:1380-1389
  117. Lang CA, Davis RA. Fish oil fatty acids impair VLDL assembly and/or secretion by cultured rat hepatocytes. J Lipid Res 1990; 31:2079-2086
  118. Jump DB, Tripathy S, Depner CM. Fatty acid-regulated transcription factors in the liver. Annu Rev Nutr 2013; 33:249-269
  119. Park Y, Harris WS. Omega-3 fatty acid supplementation accelerates chylomicron triglyceride clearance. J Lipid Res 2003; 44:455-463
  120. Davidson MH, Maki KC, Bays H, Carter R, Ballantyne CM. Effects of prescription omega-3-acid ethyl esters on lipoprotein particle concentrations, apolipoproteins AI and CIII, and lipoprotein-associated phospholipase A(2) mass in statin-treated subjects with hypertriglyceridemia. J Clin Lipidol 2009; 3:332-340
  121. Phillipson BE, Rothrock DW, Connor WE, Harris WS, Illingworth DR. Reduction of plasma lipids, lipoproteins, and apoproteins by dietary fish oils in patients with hypertriglyceridemia. N Engl J Med 1985; 312:1210-1216
  122. Skulas-Ray AC, Alaupovic P, Kris-Etherton PM, West SG. Dose-response effects of marine omega-3 fatty acids on apolipoproteins, apolipoprotein-defined lipoprotein subclasses, and Lp-PLA2 in individuals with moderate hypertriglyceridemia. J Clin Lipidol 2015; 9:360-367
  123. Sahebkar A, Simental-Mendia LE, Mikhailidis DP, Pirro M, Banach M, Sirtori CR, Reiner Z. Effect of omega-3 supplements on plasma apolipoprotein C-III concentrations: a systematic review and meta-analysis of randomized controlled trials. Ann Med 2018:1-11
  124. Davidson MH, Johnson J, Rooney MW, Kyle ML, Kling DF. A novel omega-3 free fatty acid formulation has dramatically improved bioavailability during a low-fat diet compared with omega-3-acid ethyl esters: the ECLIPSE (Epanova((R)) compared to Lovaza((R)) in a pharmacokinetic single-dose evaluation) study. J Clin Lipidol 2012; 6:573-584
  125. Offman E, Marenco T, Ferber S, Johnson J, Kling D, Curcio D, Davidson M. Steady-state bioavailability of prescription omega-3 on a low-fat diet is significantly improved with a free fatty acid formulation compared with an ethyl ester formulation: the ECLIPSE II study. Vasc Health Risk Manag 2013; 9:563-573
  126. Burr ML, Ashfield-Watt PA, Dunstan FD, Fehily AM, Breay P, Ashton T, Zotos PC, Haboubi NA, Elwood PC. Lack of benefit of dietary advice to men with angina: results of a controlled trial. Eur J Clin Nutr 2003; 57:193-200
  127. Burr ML, Fehily AM, Gilbert JF, Rogers S, Holliday RM, Sweetnam PM, Elwood PC, Deadman NM. Effects of changes in fat, fish, and fibre intakes on death and myocardial reinfarction: diet and reinfarction trial (DART). Lancet 1989; 2:757-761
  128. Dietary supplementation with n-3 polyunsaturated fatty acids and vitamin E after myocardial infarction: results of the GISSI-Prevenzione trial. Gruppo Italiano per lo Studio della Sopravvivenza nell'Infarto miocardico. Lancet1999; 354:447-455
  129. Tavazzi L, Maggioni AP, Marchioli R, Barlera S, Franzosi MG, Latini R, Lucci D, Nicolosi GL, Porcu M, Tognoni G, Gissi HFI. Effect of n-3 polyunsaturated fatty acids in patients with chronic heart failure (the GISSI-HF trial): a randomised, double-blind, placebo-controlled trial. Lancet 2008; 372:1223-1230
  130. Rauch B, Schiele R, Schneider S, Diller F, Victor N, Gohlke H, Gottwik M, Steinbeck G, Del Castillo U, Sack R, Worth H, Katus H, Spitzer W, Sabin G, Senges J, Omega Study Group. OMEGA, a randomized, placebo-controlled trial to test the effect of highly purified omega-3 fatty acids on top of modern guideline-adjusted therapy after myocardial infarction. Circulation 2010; 122:2152-2159
  131. Kromhout D, Giltay EJ, Geleijnse JM, Alpha Omega Trial Group. n-3 fatty acids and cardiovascular events after myocardial infarction. N Engl J Med 2010; 363:2015-2026
  132. Galan P, Kesse-Guyot E, Czernichow S, Briancon S, Blacher J, Hercberg S. Effects of B vitamins and omega 3 fatty acids on cardiovascular diseases: a randomised placebo controlled trial. BMJ 2010; 341:c6273
  133. Investigators OT, Bosch J, Gerstein HC, Dagenais GR, Diaz R, Dyal L, Jung H, Maggiono AP, Probstfield J, Ramachandran A, Riddle MC, Ryden LE, Yusuf S. n-3 fatty acids and cardiovascular outcomes in patients with dysglycemia. N Engl J Med 2012; 367:309-318
  134. Risk, Prevention Study Collaborative Group. n-3 fatty acids in patients with multiple cardiovascular risk factors. N Engl J Med 2013; 368:1800-1808
  135. Ascend Study Collaborative Group, Bowman L, Mafham M, Wallendszus K, Stevens W, Buck G, Barton J, Murphy K, Aung T, Haynes R, Cox J, Murawska A, Young A, Lay M, Chen F, Sammons E, Waters E, Adler A, Bodansky J, Farmer A, McPherson R, Neil A, Simpson D, Peto R, Baigent C, Collins R, Parish S, Armitage J. Effects of n-3 Fatty Acid Supplements in Diabetes Mellitus. N Engl J Med 2018; 379:1540-1550
  136. Manson JE, Cook NR, Lee IM, Christen W, Bassuk SS, Mora S, Gibson H, Albert CM, Gordon D, Copeland T, D'Agostino D, Friedenberg G, Ridge C, Bubes V, Giovannucci EL, Willett WC, Buring JE, Group VR. Marine n-3 Fatty Acids and Prevention of Cardiovascular Disease and Cancer. N Engl J Med 2019; 380: 23-32
  137. Saito Y, Yokoyama M, Origasa H, Matsuzaki M, Matsuzawa Y, Ishikawa Y, Oikawa S, Sasaki J, Hishida H, Itakura H, Kita T, Kitabatake A, Nakaya N, Sakata T, Shimada K, Shirato K, Jelis Investigators Japan. Effects of EPA on coronary artery disease in hypercholesterolemic patients with multiple risk factors: sub-analysis of primary prevention cases from the Japan EPA Lipid Intervention Study (JELIS). Atherosclerosis 2008; 200:135-140
  138. Bhatt DL, Steg PG, Miller M, Brinton EA, Jacobson TA, Ketchum SB, Doyle RT, Jr., Juliano RA, Jiao L, Granowitz C, Tardif JC, Ballantyne CM, REDUCE-IT Investigators. Cardiovascular Risk Reduction with Icosapent Ethyl for Hypertriglyceridemia. N Engl J Med 2019; 380:11-22
  139. Nicholls SJ, Lincoff AM, Garcia M, Bash D, Ballantyne CM, Barter PJ, Davidson MH, Kastelein JJP, Koenig W, McGuire DK, Mozaffarian D, Ridker PM, Ray KK, Katona BG, Himmelmann A, Loss LE, Rensfeldt M, Lundstrom T, Agrawal R, Menon V, Wolski K, Nissen SE. Effect of High-Dose Omega-3 Fatty Acids vs Corn Oil on Major Adverse Cardiovascular Events in Patients at High Cardiovascular Risk: The STRENGTH Randomized Clinical Trial. JAMA 2020; 324:2268-2280
  140. Kalstad AA, Myhre PL, Laake K, Tveit SH, Schmidt EB, Smith P, Nilsen DWT, Tveit A, Fagerland MW, Solheim S, Seljeflot I, Arnesen H, Omemi Investigators. Effects of n-3 Fatty Acid Supplements in Elderly Patients After Myocardial Infarction: A Randomized, Controlled Trial. Circulation 2021; 143:528-539
  141. Mason RP, Libby P, Bhatt DL. Emerging Mechanisms of Cardiovascular Protection for the Omega-3 Fatty Acid Eicosapentaenoic Acid. Arterioscler Thromb Vasc Biol 2020:ATVBAHA119313286
  142. Marston NA, Giugliano RP, Im K, Silverman MG, O'Donoghue ML, Wiviott SD, Ference BA, Sabatine MS. Association Between Triglyceride Lowering and Reduction of Cardiovascular Risk Across Multiple Lipid-Lowering Therapeutic Classes: A Systematic Review and Meta-Regression Analysis of Randomized Controlled Trials. Circulation 2019; 140:1308-1317
  143. Ridker PM, Rifai N, MacFadyen J, Glynn RJ, Jiao L, Steg PG, Miller M, Brinton EA, Jacobson TA, Tardif JC, Ballantyne CM, Mason RP, Bhatt DL. Effects of Randomized Treatment With Icosapent Ethyl and a Mineral Oil Comparator on Interleukin-1beta, Interleukin-6, C-Reactive Protein, Oxidized Low-Density Lipoprotein Cholesterol, Homocysteine, Lipoprotein(a), and Lipoprotein-Associated Phospholipase A2: A REDUCE-IT Biomarker Substudy. Circulation 2022; 146:372-379
  144. Goff ZD, Nissen SE. N-3 polyunsaturated fatty acids for cardiovascular risk. Curr Opin Cardiol 2022; 37:356-363
  145. Mason RP, Sherratt SCR, Eckel RH. Omega-3-fatty acids: Do they prevent cardiovascular disease? Best Pract Res Clin Endocrinol Metab 2023; 37:101681
  146. Wachira JK, Larson MK, Harris WS. n-3 Fatty acids affect haemostasis but do not increase the risk of bleeding: clinical observations and mechanistic insights. Br J Nutr 2014; 111:1652-1662
  147. Zimetbaum P, Frishman WH, Kahn S. Effects of gemfibrozil and other fibric acid derivatives on blood lipids and lipoproteins. J Clin Pharmacol 1991; 31:25-37
  148. Loomba RS, Arora R. Prevention of cardiovascular disease utilizing fibrates--a pooled meta-analysis. Am J Ther2010; 17:e182-188
  149. Rosenson RS. Fenofibrate: treatment of hyperlipidemia and beyond. Expert Rev Cardiovasc Ther 2008; 6:1319-1330
  150. Miller M, Bachorik PS, McCrindle BW, Kwiterovich PO, Jr. Effect of gemfibrozil in men with primary isolated low high-density lipoprotein cholesterol: a randomized, double-blind, placebo-controlled, crossover study. Am J Med1993; 94:7-12
  151. Berthold HK, Gouni-Berthold I. Hyperlipoproteinemia(a): clinical significance and treatment options. Atheroscler Suppl 2013; 14:1-5
  152. Frick MH, Elo O, Haapa K, Heinonen OP, Heinsalmi P, Helo P, Huttunen JK, Kaitaniemi P, Koskinen P, Manninen V, et al. Helsinki Heart Study: primary-prevention trial with gemfibrozil in middle-aged men with dyslipidemia. Safety of treatment, changes in risk factors, and incidence of coronary heart disease. N Engl J Med 1987; 317:1237-1245
  153. Rubins HB, Robins SJ, Collins D, Fye CL, Anderson JW, Elam MB, Faas FH, Linares E, Schaefer EJ, Schectman G, Wilt TJ, Wittes J. Gemfibrozil for the secondary prevention of coronary heart disease in men with low levels of high-density lipoprotein cholesterol. Veterans Affairs High-Density Lipoprotein Cholesterol Intervention Trial Study Group. N Engl J Med 1999; 341:410-418
  154. Bezafibrate Infarction Prevention s. Secondary prevention by raising HDL cholesterol and reducing triglycerides in patients with coronary artery disease. Circulation 2000; 102:21-27
  155. Czupryniak L, Joshi SR, Gogtay JA, Lopez M. Effect of micronized fenofibrate on microvascular complications of type 2 diabetes: a systematic review. Expert Opin Pharmacother 2016; 17:1463-1473
  156. Keech A, Simes RJ, Barter P, Best J, Scott R, Taskinen MR, Forder P, Pillai A, Davis T, Glasziou P, Drury P, Kesaniemi YA, Sullivan D, Hunt D, Colman P, d'Emden M, Whiting M, Ehnholm C, Laakso M, investigators Fs. Effects of long-term fenofibrate therapy on cardiovascular events in 9795 people with type 2 diabetes mellitus (the FIELD study): randomised controlled trial. Lancet 2005; 366:1849-1861
  157. Aguilar-Salinas CA, Fanghanel-Salmon G, Meza E, Montes J, Gulias-Herrero A, Sanchez L, Monterrubio-Flores EA, Gonzalez-Valdez H, Gomez Perez FJ. Ciprofibrate versus gemfibrozil in the treatment of mixed hyperlipidemias: an open-label, multicenter study. Metabolism 2001; 50:729-733
  158. Knipscheer HC, de Valois JC, van den Ende B, Wouter ten Cate J, Kastelein JJ. Ciprofibrate versus gemfibrozil in the treatment of primary hyperlipidaemia. Atherosclerosis 1996; 124 Suppl:S75-81
  159. Insua A, Massari F, Rodriguez Moncalvo JJ, Ruben Zanchetta J, Insua AM. Fenofibrate of gemfibrozil for treatment of types IIa and IIb primary hyperlipoproteinemia: a randomized, double-blind, crossover study. Endocr Pract 2002; 8:96-101
  160. Jen SL, Chen JW, Lee WL, Wang SP. Efficacy and safety of fenofibrate or gemfibrozil on serum lipid profiles in Chinese patients with type IIb hyperlipidemia: a single-blind, randomized, and cross-over study. Zhonghua Yi Xue Za Zhi (Taipei) 1997; 59:217-224
  161. Ebcioglu Z, Morgan J, Carey C, Capuzzi D. Paradoxical lowering of high-density lipoprotein cholesterol level in 2 patients receiving fenofibrate and a thiazolidinedione. Ann Intern Med 2003; 139:W80
  162. Shetty C, Balasubramani M, Capps N, Milles J, Ramachandran S. Paradoxical HDL-C reduction during rosiglitazone and fibrate treatment. Diabet Med 2007; 24:94-97
  163. Collinson PO, Hjelm CJ, Canepo-Anson R. Paradoxical high-density lipoprotein reduction induced by fenofibrate and ciprofibrate. Ann Clin Biochem 1996; 33 ( Pt 2):159-161
  164. Schofield JD, Liu Y, France MW, Sandle L, Soran H. A review of paradoxical HDL-C responses to fenofibrate, illustrated by a case report. J Clin Lipidol 2014; 8:455-459
  165. Nobecourt E, Cariou B, Lambert G, Krempf M. Severe decrease in high-density lipoprotein cholesterol with the combination of fibrates and ezetimibe: A case series. J Clin Lipidol 2017; 11:289-293
  166. Grundy SM, Vega GL, Yuan Z, Battisti WP, Brady WE, Palmisano J. Effectiveness and tolerability of simvastatin plus fenofibrate for combined hyperlipidemia (the SAFARI trial). Am J Cardiol 2005; 95:462-468
  167. Choi HD, Shin WG. Safety and efficacy of statin treatment alone and in combination with fibrates in patients with dyslipidemia: a meta-analysis. Curr Med Res Opin 2014; 30:1-10
  168. Geng Q, Ren J, Chen H, Lee C, Liang W. Adverse events following statin-fenofibrate therapy versus statin alone: a meta-analysis of randomized controlled trials. Clin Exp Pharmacol Physiol 2013; 40:219-226
  169. Kontopoulos AG, Athyros VG, Papageorgiou AA, Hatzikonstandinou HA, Mayroudi MC, Boudoulas H. Effects of simvastatin and ciprofibrate alone and in combination on lipid profile, plasma fibrinogen and low density lipoprotein particle structure and distribution in patients with familial combined hyperlipidaemia and coronary artery disease. Coron Artery Dis 1996; 7:843-850
  170. Choi HD, Shin WG, Lee JY, Kang BC. Safety and efficacy of fibrate-statin combination therapy compared to fibrate monotherapy in patients with dyslipidemia: a meta-analysis. Vascul Pharmacol 2015; 65-66:23-30
  171. Ansquer JC, Bekaert I, Guy M, Hanefeld M, Simon A. Efficacy and safety of coadministration of fenofibrate and ezetimibe compared with each as monotherapy in patients with type IIb dyslipidemia and features of the metabolic syndrome: a prospective, randomized, double-blind, three-parallel arm, multicenter, comparative study. Am J Cardiovasc Drugs 2009; 9:91-101
  172. Farnier M, Freeman MW, Macdonell G, Perevozskaya I, Davies MJ, Mitchel YB, Gumbiner B, Ezetimibe Study Group. Efficacy and safety of the coadministration of ezetimibe with fenofibrate in patients with mixed hyperlipidaemia. Eur Heart J 2005; 26:897-905
  173. Farnier M, Roth E, Gil-Extremera B, Mendez GF, Macdonell G, Hamlin C, Perevozskaya I, Davies MJ, Kush D, Mitchel YB, Ezetimibe/Simvastatin + Fenofibrate Study G. Efficacy and safety of the coadministration of ezetimibe/simvastatin with fenofibrate in patients with mixed hyperlipidemia. Am Heart J 2007; 153:335 e331-338
  174. Jones PH, Goldberg AC, Knapp HR, Kelly MT, Setze CM, Stolzenbach JC, Sleep DJ. Efficacy and safety of fenofibric acid in combination with atorvastatin and ezetimibe in patients with mixed dyslipidemia. Am Heart J2010; 160:759-766
  175. McKenney J, Jones M, Abby S. Safety and efficacy of colesevelam hydrochloride in combination with fenofibrate for the treatment of mixed hyperlipidemia. Curr Med Res Opin 2005; 21:1403-1412
  176. Weisweiler P. Low-dose colestipol plus fenofibrate: effects on plasma lipoproteins, lecithin:cholesterol acyltransferase, and postheparin lipases in familial hypercholesterolemia. Metabolism 1989; 38:271-274
  177. Curtis LD, Dickson AC, Ling KL, Betteridge J. Combination treatment with cholestyramine and bezafibrate for heterozygous familial hypercholesterolaemia. BMJ 1988; 297:173-175
  178. Series JJ, Caslake MJ, Kilday C, Cruickshank A, Demant T, Lorimer AR, Packard CJ, Shepherd J. Effect of combined therapy with bezafibrate and cholestyramine on low-density lipoprotein metabolism in type IIa hypercholesterolemia. Metabolism 1989; 38:153-158
  179. Stein EA, Heimann KW. Colestipol, clofibrate, cholestyramine and combination therapy in the treatment of familial hyperbetalipoproteinaemia. S Afr Med J 1975; 49:1252-1256
  180. Pawlak M, Lefebvre P, Staels B. Molecular mechanism of PPARalpha action and its impact on lipid metabolism, inflammation and fibrosis in non-alcoholic fatty liver disease. J Hepatol 2015; 62:720-733
  181. Staels B, Maes M, Zambon A. Fibrates and future PPARalpha agonists in the treatment of cardiovascular disease. Nat Clin Pract Cardiovasc Med 2008; 5:542-553
  182. Shah A, Rader DJ, Millar JS. The effect of PPAR-alpha agonism on apolipoprotein metabolism in humans. Atherosclerosis 2010; 210:35-40
  183. Staels B, Dallongeville J, Auwerx J, Schoonjans K, Leitersdorf E, Fruchart JC. Mechanism of action of fibrates on lipid and lipoprotein metabolism. Circulation 1998; 98:2088-2093
  184. Gaudet D, Brisson D, Tremblay K, Alexander VJ, Singleton W, Hughes SG, Geary RS, Baker BF, Graham MJ, Crooke RM, Witztum JL. Targeting APOC3 in the familial chylomicronemia syndrome. N Engl J Med 2014; 371:2200-2206
  185. Gordts PL, Nock R, Son NH, Ramms B, Lew I, Gonzales JC, Thacker BE, Basu D, Lee RG, Mullick AE, Graham MJ, Goldberg IJ, Crooke RM, Witztum JL, Esko JD. ApoC-III inhibits clearance of triglyceride-rich lipoproteins through LDL family receptors. J Clin Invest 2016; 126:2855-2866
  186. Clavey V, Lestavel-Delattre S, Copin C, Bard JM, Fruchart JC. Modulation of lipoprotein B binding to the LDL receptor by exogenous lipids and apolipoproteins CI, CII, CIII, and E. Arterioscler Thromb Vasc Biol 1995; 15:963-971
  187. Nigon F, Lesnik P, Rouis M, Chapman MJ. Discrete subspecies of human low density lipoproteins are heterogeneous in their interaction with the cellular LDL receptor. J Lipid Res 1991; 32:1741-1753
  188. A co-operative trial in the primary prevention of ischaemic heart disease using clofibrate. Report from the Committee of Principal Investigators. Br Heart J 1978; 40:1069-1118
  189. Manninen V, Elo MO, Frick MH, Haapa K, Heinonen OP, Heinsalmi P, Helo P, Huttunen JK, Kaitaniemi P, Koskinen P, et al. Lipid alterations and decline in the incidence of coronary heart disease in the Helsinki Heart Study. JAMA1988; 260:641-651
  190. Manninen V, Tenkanen L, Koskinen P, Huttunen JK, Manttari M, Heinonen OP, Frick MH. Joint effects of serum triglyceride and LDL cholesterol and HDL cholesterol concentrations on coronary heart disease risk in the Helsinki Heart Study. Implications for treatment. Circulation 1992; 85:37-45
  191. Tenkanen L, Manttari M, Manninen V. Some coronary risk factors related to the insulin resistance syndrome and treatment with gemfibrozil. Experience from the Helsinki Heart Study. Circulation 1995; 92:1779-1785
  192. Robins SJ, Collins D, Wittes JT, Papademetriou V, Deedwania PC, Schaefer EJ, McNamara JR, Kashyap ML, Hershman JM, Wexler LF, Rubins HB, Trial V-HSGVAH-DLI. Relation of gemfibrozil treatment and lipid levels with major coronary events: VA-HIT: a randomized controlled trial. JAMA 2001; 285:1585-1591
  193. Tenenbaum A, Motro M, Fisman EZ, Tanne D, Boyko V, Behar S. Bezafibrate for the secondary prevention of myocardial infarction in patients with metabolic syndrome. Arch Intern Med 2005; 165:1154-1160
  194. Meade T, Zuhrie R, Cook C, Cooper J. Bezafibrate in men with lower extremity arterial disease: randomised controlled trial. BMJ 2002; 325:1139
  195. Meade TW, For the British Medical Research Council General Practice Research Framework, participating vascular clinics. Design and intermediate results of the Lower Extremity Arterial Disease Event Reduction (LEADER)* trial of bezafibrate in men with lower extremity arterial disease [ISRCTN4119421]. Curr Control Trials Cardiovasc Med 2001; 2:195-204
  196. Scott R, O'Brien R, Fulcher G, Pardy C, D'Emden M, Tse D, Taskinen MR, Ehnholm C, Keech A, Fenofibrate Intervention and Event Lowering in Diabetes Study. Effects of fenofibrate treatment on cardiovascular disease risk in 9,795 individuals with type 2 diabetes and various components of the metabolic syndrome: the Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) study. Diabetes Care 2009; 32:493-498
  197. ACCORD Study Group, Ginsberg HN, Elam MB, Lovato LC, Crouse JR, 3rd, Leiter LA, Linz P, Friedewald WT, Buse JB, Gerstein HC, Probstfield J, Grimm RH, Ismail-Beigi F, Bigger JT, Goff DC, Jr., Cushman WC, Simons-Morton DG, Byington RP. Effects of combination lipid therapy in type 2 diabetes mellitus. N Engl J Med 2010; 362:1563-1574
  198. Das Pradhan A, Glynn RJ, Fruchart JC, MacFadyen JG, Zaharris ES, Everett BM, Campbell SE, Oshima R, Amarenco P, Blom DJ, Brinton EA, Eckel RH, Elam MB, Felicio JS, Ginsberg HN, Goudev A, Ishibashi S, Joseph J, Kodama T, Koenig W, Leiter LA, Lorenzatti AJ, Mankovsky B, Marx N, Nordestgaard BG, Pall D, Ray KK, Santos RD, Soran H, Susekov A, Tendera M, Yokote K, Paynter NP, Buring JE, Libby P, Ridker PM. Triglyceride Lowering with Pemafibrate to Reduce Cardiovascular Risk. N Engl J Med 2022; 387:1923-1934
  199. Harrold BP, Marmion VJ, Gough KR. A double-blind controlled trial of clofibrate in the treatment of diabetic retinopathy. Diabetes 1969; 18:285-291
  200. Duncan LJ, Cullen JF, Ireland JT, Nolan J, Clarke BF, Oliver MF. A three-year trial of atromid therapy in exudative diabetic retinopathy. Diabetes 1968; 17:458-467
  201. Keech AC, Mitchell P, Summanen PA, O'Day J, Davis TM, Moffitt MS, Taskinen MR, Simes RJ, Tse D, Williamson E, Merrifield A, Laatikainen LT, d'Emden MC, Crimet DC, O'Connell RL, Colman PG. Effect of fenofibrate on the need for laser treatment for diabetic retinopathy (FIELD study): a randomised controlled trial. Lancet 2007; 370:1687-1697
  202. ACCORD Study Group, ACCORD Eye Study Group, Chew EY, Ambrosius WT, Davis MD, Danis RP, Gangaputra S, Greven CM, Hubbard L, Esser BA, Lovato JF, Perdue LH, Goff DC, Jr., Cushman WC, Ginsberg HN, Elam MB, Genuth S, Gerstein HC, Schubart U, Fine LJ. Effects of medical therapies on retinopathy progression in type 2 diabetes. N Engl J Med 2010; 363:233-244
  203. Emmerich KH, Poritis N, Stelmane I, Klindzane M, Erbler H, Goldsteine J, Gortelmeyer R. [Efficacy and safety of etofibrate in patients with non-proliferative diabetic retinopathy]. Klin Monbl Augenheilkd 2009; 226:561-567
  204. Massin P, Peto T, Ansquer JC, Aubonnet P, Macu FENSIFT. Effects of fenofibric acid on diabetic macular edema: the MacuFen study. Ophthalmic Epidemiol 2014; 21:307-317
  205. Knickelbein JE, Abbott AB, Chew EY. Fenofibrate and Diabetic Retinopathy. Curr Diab Rep 2016; 16:90
  206. Hu Y, Chen Y, Ding L, He X, Takahashi Y, Gao Y, Shen W, Cheng R, Chen Q, Qi X, Boulton ME, Ma JX. Pathogenic role of diabetes-induced PPAR-alpha down-regulation in microvascular dysfunction. Proc Natl Acad Sci U S A 2013; 110:15401-15406
  207. Ansquer JC, Foucher C, Rattier S, Taskinen MR, Steiner G, DAIS Investigators. Fenofibrate reduces progression to microalbuminuria over 3 years in a placebo-controlled study in type 2 diabetes: results from the Diabetes Atherosclerosis Intervention Study (DAIS). Am J Kidney Dis 2005; 45:485-493
  208. Davis TM, Ting R, Best JD, Donoghoe MW, Drury PL, Sullivan DR, Jenkins AJ, O'Connell RL, Whiting MJ, Glasziou PP, Simes RJ, Kesaniemi YA, Gebski VJ, Scott RS, Keech ACi. Effects of fenofibrate on renal function in patients with type 2 diabetes mellitus: the Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) Study. Diabetologia 2011; 54:280-290
  209. Ansquer JC, Dalton RN, Causse E, Crimet D, Le Malicot K, Foucher C. Effect of fenofibrate on kidney function: a 6-week randomized crossover trial in healthy people. Am J Kidney Dis 2008; 51:904-913
  210. Sun X, Liu J, Wang G. Fenofibrate decreased microalbuminuria in the type 2 diabetes patients with hypertriglyceridemia. Lipids Health Dis 2020; 19:103
  211. Kouroumichakis I, Papanas N, Zarogoulidis P, Liakopoulos V, Maltezos E, Mikhailidis DP. Fibrates: therapeutic potential for diabetic nephropathy? Eur J Intern Med 2012; 23:309-316
  212. Rajamani K, Colman PG, Li LP, Best JD, Voysey M, D'Emden MC, Laakso M, Baker JR, Keech AC. Effect of fenofibrate on amputation events in people with type 2 diabetes mellitus (FIELD study): a prespecified analysis of a randomised controlled trial. Lancet 2009; 373:1780-1788
  213. Waldman B, Ansquer JC, Sullivan DR, Jenkins AJ, McGill N, Buizen L, Davis TME, Best JD, Li L, Feher MD, Foucher C, Kesaniemi YA, Flack J, d'Emden MC, Scott RS, Hedley J, Gebski V, Keech AC. Effect of fenofibrate on uric acid and gout in type 2 diabetes: a post-hoc analysis of the randomised, controlled FIELD study. Lancet Diabetes Endocrinol 2018; 6:310-318
  214. Derosa G, Maffioli P, Sahebkar A. Plasma uric acid concentrations are reduced by fenofibrate: A systematic review and meta-analysis of randomized placebo-controlled trials. Pharmacol Res 2015; 102:63-70
  215. Davidson MH, Armani A, McKenney JM, Jacobson TA. Safety considerations with fibrate therapy. Am J Cardiol2007; 99:3C-18C
  216. Sahebkar A, Simental-Mendia LE, Pirro M, Montecucco F, Carbone F, Banach M, Barreto GE, Butler AE. Impact of fibrates on circulating cystatin C levels: a systematic review and meta-analysis of clinical trials. Ann Med 2018:1-9
  217. Hadjivasilis A, Kouis P, Kousios A, Panayiotou A. The Effect of Fibrates on Kidney Function and Chronic Kidney Disease Progression: A Systematic Review and Meta-Analysis of Randomised Studies. J Clin Med 2022; 11
  218. Hottelart C, El Esper N, Rose F, Achard JM, Fournier A. Fenofibrate increases creatininemia by increasing metabolic production of creatinine. Nephron 2002; 92:536-541
  219. Hottelart C, el Esper N, Achard JM, Pruna A, Fournier A. [Fenofibrate increases blood creatinine, but does not change the glomerular filtration rate in patients with mild renal insufficiency]. Nephrologie 1999; 20:41-44
  220. National Kidney Foundation. KDOQI Clinical Practice Guideline for Diabetes and CKD: 2012 Update. Am J Kidney Dis 2012; 60:850-886
  221. Gallbladder disease as a side effect of drugs influencing lipid metabolism. Experience in the Coronary Drug Project. N Engl J Med 1977; 296:1185-1190
  222. Caroli-Bosc FX, Le Gall P, Pugliese P, Delabre B, Caroli-Bosc C, Demarquay JF, Delmont JP, Rampal P, Montet JC. Role of fibrates and HMG-CoA reductase inhibitors in gallstone formation: epidemiological study in an unselected population. Dig Dis Sci 2001; 46:540-544
  223. Hall MJ, Nelson LM, Russell RI, Howard AN. Gemfibrozil--the effect of biliary cholesterol saturation of a new lipid-lowering agent and its comparison with clofibrate. Atherosclerosis 1981; 39:511-516
  224. Preiss D, Tikkanen MJ, Welsh P, Ford I, Lovato LC, Elam MB, LaRosa JC, DeMicco DA, Colhoun HM, Goldenberg I, Murphy MJ, MacDonald TM, Pedersen TR, Keech AC, Ridker PM, Kjekshus J, Sattar N, McMurray JJ. Lipid-modifying therapies and risk of pancreatitis: a meta-analysis. JAMA 2012; 308:804-811
  225. Bonovas S, Nikolopoulos GK, Bagos PG. Use of fibrates and cancer risk: a systematic review and meta-analysis of 17 long-term randomized placebo-controlled trials. PLoS One 2012; 7:e45259
  226. Ahmad J, Odin JA, Hayashi PH, Chalasani N, Fontana RJ, Barnhart H, Cirulli ET, Kleiner DE, Hoofnagle JH. Identification and Characterization of Fenofibrate-Induced Liver Injury. Dig Dis Sci 2017; 62:3596-3604
  227. Simental-Mendia LE, Simental-Mendia M, Sanchez-Garcia A, Banach M, Atkin SL, Gotto AM, Jr., Sahebkar A. Effect of fibrates on glycemic parameters: A systematic review and meta-analysis of randomized placebo-controlled trials. Pharmacol Res 2018; 132:232-241
  228. Graham DJ, Staffa JA, Shatin D, Andrade SE, Schech SD, La Grenade L, Gurwitz JH, Chan KA, Goodman MJ, Platt R. Incidence of hospitalized rhabdomyolysis in patients treated with lipid-lowering drugs. JAMA 2004; 292:2585-2590
  229. Gaist D, Rodriguez LA, Huerta C, Hallas J, Sindrup SH. Lipid-lowering drugs and risk of myopathy: a population-based follow-up study. Epidemiology 2001; 12:565-569
  230. Alsheikh-Ali AA, Kuvin JT, Karas RH. Risk of adverse events with fibrates. Am J Cardiol 2004; 94:935-938
  231. Staffa JA, Chang J, Green L. Cerivastatin and reports of fatal rhabdomyolysis. N Engl J Med 2002; 346:539-540
  232. Jones PH, Davidson MH. Reporting rate of rhabdomyolysis with fenofibrate + statin versus gemfibrozil + any statin. Am J Cardiol 2005; 95:120-122
  233. Tobert JA. Efficacy and long-term adverse effect pattern of lovastatin. Am J Cardiol 1988; 62:28J-34J
  234. Stone NJ, Robinson JG, Lichtenstein AH, Bairey Merz CN, Blum CB, Eckel RH, Goldberg AC, Gordon D, Levy D, Lloyd-Jones DM, McBride P, Schwartz JS, Shero ST, Smith SC, Jr., Watson K, Wilson PW, Eddleman KM, Jarrett NM, LaBresh K, Nevo L, Wnek J, Anderson JL, Halperin JL, Albert NM, Bozkurt B, Brindis RG, Curtis LH, DeMets D, Hochman JS, Kovacs RJ, Ohman EM, Pressler SJ, Sellke FW, Shen WK, Smith SC, Jr., Tomaselli GF, American College of Cardiology/American Heart Association Task Force on Practice Group. 2013 ACC/AHA guideline on the treatment of blood cholesterol to reduce atherosclerotic cardiovascular risk in adults: a report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines. Circulation2014; 129:S1-45
  235. Kellick KA, Bottorff M, Toth PP, The National Lipid Association's Safety Task Force. A clinician's guide to statin drug-drug interactions. J Clin Lipidol 2014; 8:S30-46
  236. Whitfield LR, Porcari AR, Alvey C, Abel R, Bullen W, Hartman D. Effect of gemfibrozil and fenofibrate on the pharmacokinetics of atorvastatin. J Clin Pharmacol 2011; 51:378-388
  237. Prueksaritanont T, Tang C, Qiu Y, Mu L, Subramanian R, Lin JH. Effects of fibrates on metabolism of statins in human hepatocytes. Drug Metab Dispos 2002; 30:1280-1287
  238. Chait A, Subramanian S. Hypertriglyceridemia: Pathophysiology, Role of Genetics, Consequences, and Treatment. In: Feingold KR, Anawalt B, Boyce A, Chrousos G, Dungan K, Grossman A, Hershman JM, Kaltsas G, Koch C, Kopp P, Korbonits M, McLachlan R, Morley JE, New M, Perreault L, Purnell J, Rebar R, Singer F, Trence DL, Vinik A, Wilson DP, eds. Endotext. South Dartmouth (MA) 2019.
  239. Shah AS, Wilson DP. Genetic Disorders Causing Hypertriglyceridemia in Children and Adolescents. In: Feingold KR, Anawalt B, Boyce A, Chrousos G, Dungan K, Grossman A, Hershman JM, Kaltsas G, Koch C, Kopp P, Korbonits M, McLachlan R, Morley JE, New M, Perreault L, Purnell J, Rebar R, Singer F, Trence DL, Vinik A, Wilson DP, eds. Endotext. South Dartmouth (MA) 2020.
  240. Chait A, Eckel RH. The Chylomicronemia Syndrome Is Most Often Multifactorial: A Narrative Review of Causes and Treatment. Ann Intern Med 2019; 170:626-634
  241. Witztum JL, Gaudet D, Freedman SD, Alexander VJ, Digenio A, Williams KR, Yang Q, Hughes SG, Geary RS, Arca M, Stroes ESG, Bergeron J, Soran H, Civeira F, Hemphill L, Tsimikas S, Blom DJ, O'Dea L, Bruckert E. Volanesorsen and Triglyceride Levels in Familial Chylomicronemia Syndrome. N Engl J Med 2019; 381:531-542
  242. Witztum JL, Gaudet D, Arca M, Jones A, Soran H, Gouni-Berthold I, Stroes ESG, Alexander VJ, Jones R, Watts L, Xia S, Tsimikas S. Volanesorsen and triglyceride levels in familial chylomicronemia syndrome: Long-term efficacy and safety data from patients in an open-label extension trial. J Clin Lipidol 2023; 17:342-355
  243. Arca M, Hsieh A, Soran H, Rosenblit P, O'Dea L, Stevenson M. The effect of volanesorsen treatment on the burden associated with familial chylomicronemia syndrome: the results of the ReFOCUS study. Expert Rev Cardiovasc Ther 2018; 16:537-546
  244. Gaudet D, Alexander VJ, Baker BF, Brisson D, Tremblay K, Singleton W, Geary RS, Hughes SG, Viney NJ, Graham MJ, Crooke RM, Witztum JL, Brunzell JD, Kastelein JJ. Antisense Inhibition of Apolipoprotein C-III in Patients with Hypertriglyceridemia. N Engl J Med 2015; 373:438-447
  245. Gouni-Berthold I, Alexander VJ, Yang Q, Hurh E, Steinhagen-Thiessen E, Moriarty PM, Hughes SG, Gaudet D, Hegele RA, O'Dea LSL, Stroes ESG, Tsimikas S, Witztum JL. Efficacy and safety of volanesorsen in patients with multifactorial chylomicronaemia (COMPASS): a multicentre, double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Diabetes Endocrinol 2021; 9:264-275
  246. Digenio A, Dunbar RL, Alexander VJ, Hompesch M, Morrow L, Lee RG, Graham MJ, Hughes SG, Yu R, Singleton W, Baker BF, Bhanot S, Crooke RM. Antisense-Mediated Lowering of Plasma Apolipoprotein C-III by Volanesorsen Improves Dyslipidemia and Insulin Sensitivity in Type 2 Diabetes. Diabetes Care 2016; 39:1408-1415
  247. Oral EA, Garg A, Tami J, Huang EA, O'Dea LSL, Schmidt H, Tiulpakov A, Mertens A, Alexander VJ, Watts L, Hurh E, Witztum JL, Geary RS, Tsimikas S. Assessment of efficacy and safety of volanesorsen for treatment of metabolic complications in patients with familial partial lipodystrophy: Results of the BROADEN study: Volanesorsen in FPLD; The BROADEN Study. J Clin Lipidol 2022; 16:833-849
  248. Graham MJ, Lee RG, Bell TA, 3rd, Fu W, Mullick AE, Alexander VJ, Singleton W, Viney N, Geary R, Su J, Baker BF, Burkey J, Crooke ST, Crooke RM. Antisense oligonucleotide inhibition of apolipoprotein C-III reduces plasma triglycerides in rodents, nonhuman primates, and humans. Circ Res 2013; 112:1479-1490
  249. Paik J, Duggan S. Volanesorsen: First Global Approval. Drugs 2019; 79:1349-1354
  250. Hegele RA, Tsimikas S. Lipid-Lowering Agents. Circ Res 2019; 124:386-404
  251. Prohaska TA, Alexander VJ, Karwatowska-Prokopczuk E, Tami J, Xia S, Witztum JL, Tsimikas S. APOC3 inhibition with volanesorsen reduces hepatic steatosis in patients with severe hypertriglyceridemia. J Clin Lipidol 2023; 17:406-411
  252. Lee SJ, Campos H, Moye LA, Sacks FM. LDL containing apolipoprotein CIII is an independent risk factor for coronary events in diabetic patients. Arterioscler Thromb Vasc Biol 2003; 23:853-858
  253. Luc G, Fievet C, Arveiler D, Evans AE, Bard JM, Cambien F, Fruchart JC, Ducimetiere P. Apolipoproteins C-III and E in apoB- and non-apoB-containing lipoproteins in two populations at contrasting risk for myocardial infarction: the ECTIM study. Etude Cas Temoins sur 'Infarctus du Myocarde. J Lipid Res 1996; 37:508-517
  254. Sacks FM, Alaupovic P, Moye LA, Cole TG, Sussex B, Stampfer MJ, Pfeffer MA, Braunwald E. VLDL, apolipoproteins B, CIII, and E, and risk of recurrent coronary events in the Cholesterol and Recurrent Events (CARE) trial. Circulation 2000; 102:1886-1892
  255. Qamar A, Khetarpal SA, Khera AV, Qasim A, Rader DJ, Reilly MP. Plasma apolipoprotein C-III levels, triglycerides, and coronary artery calcification in type 2 diabetics. Arterioscler Thromb Vasc Biol 2015; 35:1880-1888
  256. Pollin TI, Damcott CM, Shen H, Ott SH, Shelton J, Horenstein RB, Post W, McLenithan JC, Bielak LF, Peyser PA, Mitchell BD, Miller M, O'Connell JR, Shuldiner AR. A null mutation in human APOC3 confers a favorable plasma lipid profile and apparent cardioprotection. Science 2008; 322:1702-1705
  257. Tg, Hdl Working Group of the Exome Sequencing Project NHL, Blood Institute, Crosby J, Peloso GM, Auer PL, Crosslin DR, Stitziel NO, Lange LA, Lu Y, Tang ZZ, Zhang H, Hindy G, Masca N, Stirrups K, Kanoni S, Do R, Jun G, Hu Y, Kang HM, Xue C, Goel A, Farrall M, Duga S, Merlini PA, Asselta R, Girelli D, Olivieri O, Martinelli N, Yin W, Reilly D, Speliotes E, Fox CS, Hveem K, Holmen OL, Nikpay M, Farlow DN, Assimes TL, Franceschini N, Robinson J, North KE, Martin LW, DePristo M, Gupta N, Escher SA, Jansson JH, Van Zuydam N, Palmer CN, Wareham N, Koch W, Meitinger T, Peters A, Lieb W, Erbel R, Konig IR, Kruppa J, Degenhardt F, Gottesman O, Bottinger EP, O'Donnell CJ, Psaty BM, Ballantyne CM, Abecasis G, Ordovas JM, Melander O, Watkins H, Orho-Melander M, Ardissino D, Loos RJ, McPherson R, Willer CJ, Erdmann J, Hall AS, Samani NJ, Deloukas P, Schunkert H, Wilson JG, Kooperberg C, Rich SS, Tracy RP, Lin DY, Altshuler D, Gabriel S, Nickerson DA, Jarvik GP, Cupples LA, Reiner AP, Boerwinkle E, Kathiresan S. Loss-of-function mutations in APOC3, triglycerides, and coronary disease. N Engl J Med 2014; 371:22-31
  258. Jorgensen AB, Frikke-Schmidt R, Nordestgaard BG, Tybjaerg-Hansen A. Loss-of-function mutations in APOC3 and risk of ischemic vascular disease. N Engl J Med 2014; 371:32-41
  259. Gouni-Berthold I, Schwarz J, Berthold HK. Updates in Drug Treatment of Severe Hypertriglyceridemia. Curr Atheroscler Rep 2023; 25:701-709
  260. Scott LJ. Alipogene tiparvovec: a review of its use in adults with familial lipoprotein lipase deficiency. Drugs 2015; 75:175-182
  261. Fisher RM, Humphries SE, Talmud PJ. Common variation in the lipoprotein lipase gene: effects on plasma lipids and risk of atherosclerosis. Atherosclerosis 1997; 135:145-159
  262. Rip J, Nierman MC, Ross CJ, Jukema JW, Hayden MR, Kastelein JJ, Stroes ES, Kuivenhoven JA. Lipoprotein lipase S447X: a naturally occurring gain-of-function mutation. Arterioscler Thromb Vasc Biol 2006; 26:1236-1245
  263. Gaudet D, Methot J, Dery S, Brisson D, Essiembre C, Tremblay G, Tremblay K, de Wal J, Twisk J, van den Bulk N, Sier-Ferreira V, van Deventer S. Efficacy and long-term safety of alipogene tiparvovec (AAV1-LPLS447X) gene therapy for lipoprotein lipase deficiency: an open-label trial. Gene Ther 2013; 20:361-369
  264. Carpentier AC, Frisch F, Labbe SM, Gagnon R, de Wal J, Greentree S, Petry H, Twisk J, Brisson D, Gaudet D. Effect of alipogene tiparvovec (AAV1-LPL(S447X)) on postprandial chylomicron metabolism in lipoprotein lipase-deficient patients. J Clin Endocrinol Metab 2012; 97:1635-1644
  265. Gaudet D, Stroes ES, Methot J, Brisson D, Tremblay K, Bernelot Moens SJ, Iotti G, Rastelletti I, Ardigo D, Corzo D, Meyer C, Andersen M, Ruszniewski P, Deakin M, Bruno MJ. Long-Term Retrospective Analysis of Gene Therapy with Alipogene Tiparvovec and Its Effect on Lipoprotein Lipase Deficiency-Induced Pancreatitis. Hum Gene Ther 2016; 27:916-925
  266. Kassner U, Hollstein T, Grenkowitz T, Wuhle-Demuth M, Salewsky B, Demuth I, Dippel M, Steinhagen-Thiessen E. Gene Therapy in Lipoprotein Lipase Deficiency: Case Report on the First Patient Treated with Alipogene Tiparvovec Under Daily Practice Conditions. Hum Gene Ther 2018; 29:520-527
  267. Ahmad Z, Banerjee P, Hamon S, Chan KC, Bouzelmat A, Sasiela WJ, Pordy R, Mellis S, Dansky H, Gipe DA, Dunbar RL. Inhibition of Angiopoietin-Like Protein 3 With a Monoclonal Antibody Reduces Triglycerides in Hypertriglyceridemia. Circulation 2019; 140:470-486
  268. Rosenson RS, Gaudet D, Ballantyne CM, Baum SJ, Bergeron J, Kershaw EE, Moriarty PM, Rubba P, Whitcomb DC, Banerjee P, Gewitz A, Gonzaga-Jauregui C, McGinniss J, Ponda MP, Pordy R, Zhao J, Rader DJ. Evinacumab in severe hypertriglyceridemia with or without lipoprotein lipase pathway mutations: a phase 2 randomized trial. Nat Med 2023; 29:729-737
  269. Mattijssen F, Kersten S. Regulation of triglyceride metabolism by Angiopoietin-like proteins. Biochim Biophys Acta2012; 1821:782-789
  270. Kersten S. New insights into angiopoietin-like proteins in lipid metabolism and cardiovascular disease risk. Curr Opin Lipidol 2019; 30:205-211
  271. Musunuru K, Pirruccello JP, Do R, Peloso GM, Guiducci C, Sougnez C, Garimella KV, Fisher S, Abreu J, Barry AJ, Fennell T, Banks E, Ambrogio L, Cibulskis K, Kernytsky A, Gonzalez E, Rudzicz N, Engert JC, DePristo MA, Daly MJ, Cohen JC, Hobbs HH, Altshuler D, Schonfeld G, Gabriel SB, Yue P, Kathiresan S. Exome sequencing, ANGPTL3 mutations, and familial combined hypolipidemia. N Engl J Med 2010; 363:2220-2227
  272. Dewey FE, Gusarova V, Dunbar RL, O'Dushlaine C, Schurmann C, Gottesman O, McCarthy S, Van Hout CV, Bruse S, Dansky HM, Leader JB, Murray MF, Ritchie MD, Kirchner HL, Habegger L, Lopez A, Penn J, Zhao A, Shao W, Stahl N, Murphy AJ, Hamon S, Bouzelmat A, Zhang R, Shumel B, Pordy R, Gipe D, Herman GA, Sheu WHH, Lee IT, Liang KW, Guo X, Rotter JI, Chen YI, Kraus WE, Shah SH, Damrauer S, Small A, Rader DJ, Wulff AB, Nordestgaard BG, Tybjaerg-Hansen A, van den Hoek AM, Princen HMG, Ledbetter DH, Carey DJ, Overton JD, Reid JG, Sasiela WJ, Banerjee P, Shuldiner AR, Borecki IB, Teslovich TM, Yancopoulos GD, Mellis SJ, Gromada J, Baras A. Genetic and Pharmacologic Inactivation of ANGPTL3 and Cardiovascular Disease. N Engl J Med 2017; 377:211-221
  273. Stitziel NO, Khera AV, Wang X, Bierhals AJ, Vourakis AC, Sperry AE, Natarajan P, Klarin D, Emdin CA, Zekavat SM, Nomura A, Erdmann J, Schunkert H, Samani NJ, Kraus WE, Shah SH, Yu B, Boerwinkle E, Rader DJ, Gupta N, Frossard PM, Rasheed A, Danesh J, Lander ES, Gabriel S, Saleheen D, Musunuru K, Kathiresan S. ANGPTL3 Deficiency and Protection Against Coronary Artery Disease. J Am Coll Cardiol 2017; 69:2054-2063
  274. Berglund L, Brunzell JD, Goldberg AC, Goldberg IJ, Sacks F, Murad MH, Stalenhoef AF, Endocrine Society. Evaluation and treatment of hypertriglyceridemia: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab 2012; 97:2969-2989
  275. Jacobson TA, Ito MK, Maki KC, Orringer CE, Bays HE, Jones PH, McKenney JM, Grundy SM, Gill EA, Wild RA, Wilson DP, Brown WV. National lipid association recommendations for patient-centered management of dyslipidemia: part 1--full report. J Clin Lipidol 2015; 9:129-169
  276. Chait A, Feingold KR. Approach to patients with hypertriglyceridemia. Best Pract Res Clin Endocrinol Metab 2023; 37:101659
  277. Feingold KR. Dyslipidemia in Patients with Diabetes. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  278. Feingold KR. The Effect of Endocrine Disorders on Lipids and Lipoproteins. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.

Introduction to Lipids and Lipoproteins

ABSTRACT

 

Cholesterol and triglycerides are insoluble in water and therefore these lipids must be transported in association with proteins. Lipoproteins are complex particles with a central core containing cholesterol esters and triglycerides surrounded by free cholesterol, phospholipids, and apolipoproteins, which facilitate lipoprotein formation and function. Plasma lipoproteins can be divided into seven classes based on size, lipid composition, and apolipoproteins (chylomicrons, chylomicron remnants, VLDL, VLDL remnants (IDL), LDL, HDL, and Lp (a)).  Chylomicron remnants, VLDL, IDL, LDL, and Lp (a) are all pro-atherogenic while HDL is anti-atherogenic. Apolipoproteins have four major functions including 1) serving a structural role, 2) acting as ligands for lipoprotein receptors, 3) guiding the formation of lipoproteins, and 4) serving as activators or inhibitors of enzymes involved in the metabolism of lipoproteins. The exogenous lipoprotein pathway starts with the incorporation of dietary lipids into chylomicrons in the intestine. In the circulation, the triglycerides carried in chylomicrons are metabolized in muscle and adipose tissue by lipoprotein lipase releasing free fatty acids, which are subsequently metabolized by muscle and adipose tissue, and chylomicron remnants are formed. Chylomicron remnants are then taken up by the liver. The endogenous lipoprotein pathway begins in the liver with the formation of VLDL. The triglycerides carried in VLDL are metabolized in muscle and adipose tissue by lipoprotein lipase releasing free fatty acids and IDL are formed. The IDL are further metabolized to LDL, which are taken up by the LDL receptor in numerous tissues including the liver, the predominant site of uptake. Reverse cholesterol transport begins with the formation of nascent HDL by the liver and intestine. These small HDL particles can then acquire cholesterol and phospholipids that are effluxed from cells, a process mediated by ABCA1 resulting in the formation of mature HDL.  Mature HDL can acquire addition cholesterol from cells via ABCG1, SR-B1, or passive diffusion. The HDL then transports the cholesterol to the liver either directly by interacting with hepatic SR-B1 or indirectly by transferring the cholesterol to VLDL or LDL, a process facilitated by CETP. Cholesterol efflux from macrophages to HDL plays an important role in protecting from the development of atherosclerosis.

 

INTRODUCTION

 

Because lipids, such as cholesterol and triglycerides, are insoluble in water these lipids must be transported in association with proteins (lipoproteins) in the circulation. Large quantities of fatty acids from meals must be transported as triglycerides to avoid toxicity. These lipoproteins play a key role in the absorption and transport of dietary lipids by the small intestine, in the transport of lipids from the liver to peripheral tissues, and the transport of lipids from peripheral tissues to the liver and intestine (reverse cholesterol transport). A secondary function is to transport toxic foreign hydrophobic and amphipathic compounds, such as bacterial toxins, from areas of invasion and infection (1). For example, lipoproteins bind endotoxin (LPS) from gram negative bacteria and lipoteichoic acid from gram positive bacteria thereby reducing their toxic effects (1). In addition, apolipoprotein L1, associated with HDL particles, has lytic activity against the parasite Trypanosoma brucei brucei and lipoproteins can neutralize viruses (2,3). Thus, while this article will focus on the transport properties of lipoproteins the reader should recognize that lipoprotein may have other important roles.

 

STRUCTURE OF LIPOPROTEINS (4)

 

Lipoproteins are complex particles that have a central hydrophobic core of non-polar lipids, primarily cholesterol esters and triglycerides. This hydrophobic core is surrounded by a hydrophilic membrane consisting of phospholipids, free cholesterol, and apolipoproteins (Figure 1). Plasma lipoproteins are divided into seven classes based on size, lipid composition, and apolipoproteins (Table 1 and Figure 2).

 

Figure 1. Lipoprotein Structure (figure modified from Biochemistry 39: 9763, 2000)

 

 

Table 1. Lipoprotein Classes

Lipoprotein

Density (g/ml)

Size (nm)

Major Lipids

Major Apoproteins

Chylomicrons

<0.930

75-1200

Triglycerides

Apo B-48, Apo C, Apo E, Apo A-I, A-II, A-IV

Chylomicron Remnants

0.930- 1.006

30-80

Triglycerides Cholesterol

Apo B-48, Apo E

VLDL

0.930- 1.006

30-80

Triglycerides

Apo B-100, Apo E, Apo C

IDL

1.006- 1.019

25-35

Triglycerides Cholesterol

Apo B-100, Apo E, Apo C

LDL

1.019- 1.063

18- 25

Cholesterol

Apo B-100

HDL

1.063- 1.210

5- 12

Cholesterol Phospholipids

Apo A-I, Apo A-II, Apo C, Apo E

Lp (a)

1.055- 1.085

~30

Cholesterol

Apo B-100, Apo (a)

 

Figure 2: Classes of Lipoproteins (figure modified from Advances Protein Chemistry 45:303, 1994)

 

Chylomicrons (5)

 

These are large triglyceride rich particles made by the intestine, which are involved in the transport of dietary triglycerides and cholesterol to peripheral tissues and liver. These particles contain apolipoproteins A-I, A-II, A-IV, A-V, B-48, C-II, C-III, and E. Apo B-48 is the core structural protein and each chylomicron particle contains one Apo B-48 molecule. The size of chylomicrons varies depending on the amount of fat ingested. A high fat meal leads to the formation of large chylomicron particles due to the increased amount of triglyceride being transported whereas in the fasting state the chylomicron particles are small carrying decreased quantities of triglyceride. The quantity of cholesterol carried by chylomicrons also can vary depending upon dietary intake.

 

Chylomicron Remnants (5-7)

 

The removal of triglyceride from chylomicrons by lipoprotein lipase in peripheral tissues results in smaller particles called chylomicron remnants. Compared to chylomicrons these particles are enriched in cholesterol and are pro-atherogenic.

 

Very Low-Density Lipoproteins (VLDL)

 

These particles are produced by the liver and are triglyceride rich. They contain apolipoprotein B-100, C-I, C-II, C-III, and E. Apo B-100 is the core structural protein and each VLDL particle contains one Apo B-100 molecule. Similar to chylomicrons the size of the VLDL particles can vary depending on the quantity of triglyceride carried in the particle. When triglyceride production in the liver is increased, the secreted VLDL particles are large. However, VLDL particles are smaller than chylomicrons.

 

Intermediate-Density Lipoproteins (IDL; VLDL Remnants) (6,7)

 

The removal of triglycerides from VLDL by muscle and adipose tissue results in the formation of IDL particles which are enriched in cholesterol. These particles contain apolipoprotein B-100 and E. These IDL particles are pro-atherogenic.

 

Low-Density Lipoproteins (LDL)

 

These particles are derived from VLDL and IDL particles and they are even further enriched in cholesterol. LDL carries the majority of the cholesterol that is in the circulation. The predominant apolipoprotein is B-100 and each LDL particle contains one Apo B-100 molecule. LDL consists of a spectrum of particles varying in size and density. An abundance of small dense LDL particles is seen in association with hypertriglyceridemia, low HDL levels, obesity, type 2 diabetes (i.e. patients with the metabolic syndrome) and infectious and inflammatory states. These small dense LDL particles are considered to be more pro-atherogenic than large LDL particles for a number of reasons (8). Small dense LDL particles have a decreased affinity for the LDL receptor resulting in a prolonged retention time in the circulation. Additionally, they more easily enter the arterial wall and bind more avidly to intra-arterial proteoglycans, which traps them in the arterial wall. Finally, small dense LDL particles are more susceptible to oxidation, which could result in an enhanced uptake by macrophages. 

 

High-Density Lipoproteins (HDL) (9-11)

 

These particles play an important role in reverse cholesterol transport from peripheral tissues to the liver, which is one potential mechanism by which HDL may be anti-atherogenic. In addition, HDL particles have anti-oxidant, anti-inflammatory, anti-thrombotic, and anti-apoptotic properties, which may also contribute to their ability to inhibit atherosclerosis. HDL particles are enriched in cholesterol and phospholipids. Apolipoproteins A-I, A-II, A-IV, C-I, C-II, C-III, and E are associated with these particles. Apo A-I is the core structural protein and each HDL particle may contain multiple Apo A-I molecules. In addition, using mass spectrometry proteins involved in proteinase inhibition, complement activation, and the acute-phase response have been found associated with HDL particles (12). HDL particles are very heterogeneous and can be classified based on density, size, charge, or apolipoprotein composition (Table 2).

 

Table 2. Classification of HDL

Method of classification

Types of HDL

Density gradient ultracentrifugation

HDL2, HDL3, very high-density HDL

Nuclear magnetic resonance

large, medium, and small

Gradient gel electrophoresis

HDL 2a, 2b, 3a, 3b, 3c

2-dimensional gel electrophoresis

pre-beta 1 and 2, alpha 1, 2, 3, 4

Apolipoprotein composition

A-I particles, A-I: A-II particles, A-I: E particles

 

Lipoprotein (a) (Lp (a)) (13-16)

 

Lp (a) is an LDL particle that has apolipoprotein (a) attached to Apo B-100 via a disulfide bond. Lp (a) contain Apo (a) and Apo B-100 in a 1:1 molar ratio. The size of Lp(a) particles can vary greatly based on the size of apolipoprotein (a). This particle is pro-atherogenic.

 

APOLIPOPROTEINS (17,18)

 

Apolipoproteins have four major functions including 1) serving a structural role, 2) acting as ligands for lipoprotein receptors, 3) guiding the formation of lipoproteins, and 4) serving as activators or inhibitors of enzymes involved in the metabolism of lipoproteins (Table 3). Apolipoproteins thus play a crucial role in lipoprotein metabolism.

 

Apolipoprotein A-I (19)

 

Apo A-I is synthesized in the liver and intestine and is the major structural protein of HDL accounting for approximately 70% of HDL protein. It also plays a role in the interaction of HDL with ATP-binding cassette protein A1 (ABCA1), ABCG1, and class B, type I scavenger receptor (SR-B1). Apo A-I is an activator of lecithin: cholesterol acyltransferase (LCAT), an enzyme that converts free cholesterol into cholesteryl ester. High levels of Apo A-I are associated with a decreased risk of atherosclerosis.

 

Apolipoprotein A-II (20)

 

Apo A-II is synthesized in the liver and is the second most abundant protein on HDL accounting for approximately 20% of HDL protein. The role of Apo A-II in lipid metabolism is unclear. Apo A-II is a strong predictor of risk for CVD.

 

Apolipoprotein A-IV (21)

 

Apo A-IV is synthesized in the intestine during fat absorption. Apo A-IV is associated with chylomicrons and high-density lipoproteins, but is also found in the lipoprotein-free fraction.  Its precise role in lipoprotein metabolism remains to be determined but studies have suggested a role for Apo A-IV in regulating food intake.

 

Apolipoprotein A-V (22,23)

 

Apo A-V is synthesized in the liver and associates with triglyceride rich lipoproteins. It is an activator of LPL mediated lipolysis and thereby plays an important role in the metabolism of triglyceride rich lipoproteins.

 

Apolipoprotein B-48 (24)

 

Apo B-48 is synthesized in the intestine and is the major structural protein of chylomicrons and chylomicron remnants. There is a single molecule of apo B-48 per chylomicron particle. There is a single apolipoprotein B gene that is expressed in both the liver and intestine. The intestine expresses a protein that is approximately ½ the size of the liver due to mRNA editing. The apobec-1 editing complex is expressed in the intestine and edits a specific cytidine to an uracil in the apo B mRNA in the intestine creating a stop codon that results in the cessation of protein translation and a shorter Apo B (Apo B-48). The portion of Apo-B that is recognized by the LDL receptor is not contained in Apo-B48 and therefore Apo B-48 is not recognized by the LDL receptor.

 

Apolipoprotein B-100

 

Apo B-100 is synthesized in the liver and is the major structural component of VLDL, IDL, and LDL. There is a single molecule of Apo B-100 per VLDL, IDL, LDL and Lp(a) particle. Apo B-100 is a ligand for the LDL receptor and therefore plays an important role in the clearance of lipoprotein particles. Certain mutations in Apo B-100 result in decreased binding to the LDL receptor and familial hypercholesterolemia (25). High levels of Apo B-100 are associated with an increased risk of atherosclerosis.

 

Apolipoprotein C (26-29)

 

The C apolipoproteins are synthesized primarily in the liver and freely exchange between lipoprotein particles and therefore are found in association with chylomicrons, VLDL, and HDL.

 

Apo C-II is a co-factor for lipoprotein lipase (LPL) and thus stimulates triglyceride hydrolysis and the clearance of triglyceride rich lipoproteins (26,29). Loss of function mutations in Apo C-II result in marked hypertriglyceridemia due to a failure to metabolize triglyceride rich lipoproteins (30).

 

Apo C-III is an inhibitor of LPL (31). Additionally, Apo C-III inhibits the interaction of triglyceride rich lipoproteins with their receptors (31). Recent studies have shown that loss of function mutations in Apo C-III lead to decreases in serum triglyceride levels and a reduced risk of cardiovascular disease. Interestingly, inhibition of Apo C-III expression results in a decrease in serum triglyceride levels even in patients deficient in lipoprotein lipase indicating that the ability of Apo C-III to modulate serum triglyceride levels is not dependent solely on regulating lipoprotein lipase activity (32).

 

Apolipoprotein E (33)

 

Apolipoprotein E is synthesized in many tissues but the liver and intestine are the primary source of circulating Apo E. Apo E exchanges between lipoprotein particles and is associated with chylomicrons, chylomicron remnants, VLDL, IDL, and a subgroup of HDL particles. There are three common genetic variants of Apo E (Apo E2, E3, and E4). ApoE2 differs from the most common isoform, Apo E3, by a single amino acid substitution where cysteine substitutes for arginine at residue 158. Apo E4 differs from Apo E3 at residue 112, where arginine substitutes for cysteine. Apo E3 and E4 are ligands for the LDL receptor while Apo E2 is poorly recognized by the LDL receptor. Patients who are homozygous for Apo E2 can develop familial dysbetalipoproteinemia (30). Apo E4 is associated with an increased risk of Alzheimer’s disease and an increased risk of atherosclerosis.

 

Apolipoprotein (a) (14,16)

 

Apo (a) is synthesized in the liver. This protein is a homolog of plasminogen and its molecular weight varies from 300,000 to 800,000. It is attached to Apo B-100 via a disulfide bond. High levels of Apo (a) are associated with an increased risk of atherosclerosis. Apo (a) is an inhibitor of fibrinolysis and can also enhance the uptake of lipoproteins by macrophages, both of which could increase the risk of atherosclerosis. The physiologic function of Apo (a) is unknown. Interestingly this apolipoprotein is found in primates but not in other species.

 

Table 3. Apolipoproteins

Apolipoprotein

MW

Primary Source

Lipoprotein Association

Function

Apo A-I

28,000

Liver, Intestine

HDL, chylomicrons

Structural protein for HDL, Activates LCAT

Apo A-II

17,000

Liver

HDL, chylomicrons

Structural protein for HDL, Activates hepatic lipase

Apo A-IV

45,000

Intestine

HDL, chylomicrons

Unknown

Apo A-V

39,000

Liver

VLDL, chylomicrons, HDL               

Promotes LPL mediated TG lipolysis

Apo B-48

241,000

Intestine

Chylomicrons

Structural protein for chylomicrons

Apo B-100

512,000

Liver

VLDL, IDL, LDL, Lp (a)

Structural protein, Ligand for LDL receptor

Apo C-I

6,600

Liver

Chylomicrons, VLDL, HDL

Activates LCAT

Apo C-II

8,800

Liver

Chylomicrons, VLDL, HDL

Co-factor for LPL

Apo C-III

8,800

Liver

Chylomicrons, VLDL, HDL

Inhibits LPL and uptake of lipoproteins

Apo E

34,000

Liver

Chylomicron remnants, IDL, HDL

Ligand for LDL receptor

Apo (a)

250,000- 800,00

Liver

Lp (a)

Inhibits plasminogen activation

                                

LIPOPROTEIN RECEPTORS AND LIPID TRANSPORTERS

 

There are several receptors and transporters that play a crucial role in lipoprotein metabolism.

 

LDL Receptor (34)

 

The LDL receptor is present in the liver and most other tissues. It recognizes Apo B-100 and Apo E and hence mediates the uptake of LDL, chylomicron remnants, and IDL, which occurs via endocytosis (Figure 3). After internalization, the lipoprotein particle is degraded in lysosomes and the cholesterol is released. The delivery of cholesterol to the cell decreases the activity of HMGCoA reductase and other enzymes required for the biosynthesis of cholesterol, and the expression of LDL receptors. LDL receptors in the liver play a major role in determining plasma LDL levels (a low number of receptors is associated with high plasma LDL levels while a high number of hepatic LDL receptors is associated with low plasma LDL levels). The number of LDL receptors is regulated by the cholesterol content of the cell (35). When cellular cholesterol levels are decreased the transcription factor SREBP is transported from the endoplasmic reticulum to the Golgi where proteases cleave and activate SREBP, which then migrates to the nucleus and stimulates the expression of LDL receptors (Figure 4). Conversely, when cellular cholesterol levels are high SREBP remains in the endoplasmic reticulum in an inactive form and the expression of LDL receptors is low. As discussed later PCSK9 regulates the rate of degradation of LDL receptors.

 

Figure 3. LDL Receptor Pathway (figure modified from Annual Review of Biochemistry 46: 897, 1977)

Figure 4. SREBP Pathway (figure modified from Journal of Lipid Research 50: Supp S15, 2009)

 

 

LRP-1 is a member of the LDL receptor family. It is expressed in multiple tissues including the liver. LRP-1 recognizes Apo E and mediates the uptake of chylomicron remnants and IDL (VLDL remnants).

 

VLDL Receptor (37)

 

The VLDL receptor is a member of the LDL receptor family. The VLDLR is expressed in the heart, skeletal muscle, adipose tissue, endothelium, brain, macrophages, and other tissues. Interestingly it is not usually expressed in the liver but hepatic expression can be induced by endoplasmic reticulum stress and PPAR alpha activation. Apo E but not Apo B bind to the VLDL receptor thereby allowing for the uptake of triglyceride rich lipoprotein particles (VLDL and chylomicrons).

 

Class B Scavenger Receptor B1 (SR-B1) (38)

 

SR-B1 is expressed in the liver, adrenal glands, ovaries, testes, macrophages, and other cells. In the liver and steroid producing cells, it mediates the selective uptake of cholesterol esters from HDL particles. In macrophages and other cells, it facilitates the efflux of cholesterol from the cell to HDL particles.

 

ATP-Binding Cassette Transporter A1 (ABCA1) (39)

 

ABCA1 is expressed in many cells including hepatocytes, enterocytes, and macrophages. It mediates the transport of cholesterol and phospholipids from the cell to lipid poor HDL particles (pre-beta-HDL).

 

ATP-Binding Cassette Transporter G1 (ABCG1) (40)

 

ABCG1 is expressed in many different cell types and mediates the efflux of cholesterol from the cell to HDL particles.

 

ATP-Binding Cassette Transporter G5 and G8 (ABCG5/ABCG8) (41,42)

 

ABCG5 and ABCG8 are expressed in the liver and intestine and form a heterodimer. In the intestine, these transporters mediate the movement of plant sterols and cholesterol from inside the enterocyte into the intestinal lumen thereby decreasing the absorption of cholesterol and limiting the uptake of dietary plant sterols. In the liver, these transporters play a role in the movement of cholesterol and plant sterols into the bile facilitating the excretion of sterols.

 

Niemann-Pick C1-Like 1 (NPC1L1) (41)

 

NPC1L1 is expressed in the intestine and mediates the uptake of cholesterol and plant sterols from the intestinal lumen into the enterocyte. NPC1L1 is also expressed in the liver where it mediates the movement of cholesterol from hepatocytes into the bile.

 

ENZYMES AND TRANSFER PROTEINS INVOLVED IN LIPOPROTEIN METABOLISM

 

There are several enzymes and transfer proteins that play a key role in lipoprotein metabolism.

 

Lipoprotein Lipase (LPL) (43)

 

LPL is synthesized in muscle, heart, and adipose tissue, then secreted and attached to the endothelium of the adjacent blood capillaries. This enzyme hydrolyzes the triglycerides carried in chylomicrons and VLDL to fatty acids, which can be taken up by cells. The catabolism of triglycerides results in the conversion of chylomicrons into chylomicron remnants and VLDL into IDL (VLDL remnants). This enzyme requires Apo C-II as a cofactor. Apo A-V also plays a key role in the activation of this enzyme. In contrast Apo C-III and Apo A-II inhibit the activity of LPL. Insulin stimulates LPL expression and LPL activity is reduced in patients with poorly controlled diabetes, which can impair the metabolism of triglyceride rich lipoproteins leading to hypertriglyceridemia (44).

 

Hepatic Lipase (45)

 

Hepatic lipase is localized to the sinusoidal surface of liver cells. It mediates the hydrolysis of triglycerides and phospholipids in IDL and LDL leading to smaller particles (IDL is converted to LDL; LDL is converted from large LDL to small LDL). It also mediates the hydrolysis of triglycerides and phospholipids in HDL resulting in smaller HDL particles.

 

Endothelial Lipase (46)

 

Endothelial lipase plays a major role in hydrolyzing the phospholipids in HDL.

 

Lecithin: Cholesterol Acyltransferase (LCAT) (47)

 

LCAT is made in the liver. In the plasma, it catalyzes the synthesis of cholesterol esters in HDL by facilitating the transfer of a fatty acid from position 2 of lecithin to cholesterol. This allows for the transfer of the cholesterol from the surface of the HDL particle (free cholesterol) to the core of the HDL particle (cholesterol ester), which facilitates the continued uptake of free cholesterol by HDL particles by reducing the concentration of cholesterol on the surface of HDL.

 

Cholesteryl Ester Transfer Protein (CETP) (48,49)

 

This protein is synthesized in the liver and in the plasma mediates the transfer of cholesterol esters from HDL to VLDL, chylomicrons, and LDL and the transfer of triglycerides from VLDL and chylomicrons to HDL. Inhibition of CETP activity leads to an increase in HDL cholesterol and a decrease in LDL cholesterol.

 

Microsomal Triglyceride Transfer Protein (MTTP) (50)

 

MTTP is expressed primarily in the liver and small intestine and plays a crucial role in the synthesis of lipoproteins in these tissues. MTTP mediates the transfer of triglycerides to apolipoprotein B-100 in the liver to form VLDL and to apolipoprotein B-48 in the intestine to form chylomicrons.

 

EXOGENOUS LIPOPROTEIN PATHWAY (CHYLOMICRONS)

 

Figure 5. Exogenous Lipoprotein Pathway

 

Fat Absorption (51-54)

 

The exogenous lipoprotein pathway starts in the intestine. Dietary triglycerides (approximately 100 grams per day) are hydrolyzed to free fatty acids and monoacylglycerol by intestinal lipases and emulsified with bile acids, cholesterol, plant sterols, and fat-soluble vitamins to form micelles. While the fatty acids in the intestine are overwhelmingly accounted for by dietary intake the cholesterol in the intestinal lumen is primarily derived from bile (approximately 800-1200mg of cholesterol from bile vs. 200-500mg from diet). Plant sterols account for approximately 25% of dietary sterol intake (approximately 100-150mg/day). The cholesterol, plant sterols, fatty acids, monoacylglycerol, and fat-soluble vitamins contained in the micelles are then transported into the intestinal cells. The uptake of cholesterol and plant sterols from the intestinal lumen into intestinal cells is facilitated by a sterol transporter, Niemann-Pick C1- like 1 protein (NPC1L1) (Figure 6). Ezetimibe, a drug which inhibits intestinal cholesterol and plant sterol uptake, binds to NPC1L1 and inhibits its activity. Once in the intestinal cell the cholesterol and plant sterols may be transported back into the intestinal lumen, a process mediated by ABCG5 and ABCG8, or converted to sterol esters by acyl-CoA cholesterol acyl transferase (ACAT), which attaches a fatty acid to the sterol. Compared to cholesterol, plant sterols are poor substrates for ACAT and therefore the formation of plant sterol esters does not occur as efficiently as the formation of cholesterol esters. In humans, <5% of dietary plant sterols are absorbed and the vast majority are transported out of the intestine cell, a process mediated by ABCG5 and ABCG8, which are very efficient at effluxing plant sterols from the intestinal cell into the intestinal lumen. Patients with sitosterolemia have mutations in either ABCG5 or ABCG8 and net absorption of dietary plant sterols is increased (20-30% absorbed vs. < 5% in normal subjects) (55). Thus, ABCG5 and ABCG8 along with ACAT serve as gate keepers and block the uptake of plant sterols and likely also play an important role in determining the efficiency of cholesterol absorption (humans typically absorb only approximately 50% of dietary cholesterol with a range of 25-75%).

 

Figure 6. Intestinal Cell and Sterol Metabolism. C= cholesterol, CE= cholesterol ester.

 

The pathway of absorption of free fatty acids is not well understood but it is likely that both passive diffusion and specific transporters play a role. The fatty acid transporter CD36 is strongly expressed in the proximal third of the intestine and is localized to the villi. While this transporter likely plays a role in fatty acid uptake by intestinal cells, this transporter is not essential as humans and mice deficient in this protein do not have fat malabsorption. However, in mice deficient in CD36 there is a shift in the absorption of lipid to the distal intestine, suggesting pathways that can compensate for the absence of CD36. Fatty acid transport protein 4 (FATP4) is also highly expressed in the intestine. However, mice deficient in FATP4 do not have abnormalities in fat absorption. It is likely that there are multiple pathways for the absorption of fatty acids into intestinal cells. The pathways by which monoacylglycerols are absorbed by intestinal cells remain to be defined.

 

Formation of Chylomicrons (51,54)

 

The absorbed fatty acids and monoacylglycerols are utilized to synthesize triglycerides. The key enzymes required for triglyceride synthesis are monoacylglycerol acyltransferase (MGAT) and diacylglycerol transferase (DGAT). MGAT catalyzes the addition of a fatty acid to monoacylglycerol while DGAT catalyzes the addition of a fatty acid to diacylglycerol resulting in triglyceride formation.  As noted above, the majority of the cholesterol absorbed by the intestine is esterified to cholesterol esters by ACAT. The triglycerides and cholesterol esters are packaged into chylomicrons in the endoplasmic reticulum. The size and composition of the chylomicrons formed in the intestine are dependent on the amount of fat ingested and absorbed by the intestine and the type of fat absorbed. Increased fat absorption results in larger chylomicrons. The formation of chylomicrons in the endoplasmic reticulum requires the synthesis of Apo B-48 by the intestinal cell (Figure 6).  Microsomal triglyceride transfer protein (MTTP) is required for the movement of lipid from the endoplasmic reticulum to the Apo B-48. The absence of MTTP results in the inability to form chylomicrons (Abetalipoproteinemia) (56). Lomitapide inhibits MTTP function and is used to treat patients with homozygous Familial Hypercholesterolemia (57).

 

Chylomicron Metabolism (26,31,43,58-62)

 

Chylomicrons are secreted into the lymph and delivered via the thoracic duct to the circulation. It should be noted that this results in the newly formed chylomicrons being delivered to the systemic circulation and not delivered directly to the liver via the portal circulation. This facilitates the delivery of the nutrients contained in the chylomicrons to muscle and adipose tissue. In muscle and adipose tissue lipoprotein lipase (LPL) is expressed at high levels. LPL is synthesized in muscle and adipocytes and transported to the luminal surface of capillaries. Lipase maturation factor 1 plays a key role in the stabilization and movement of LPL from muscle cells and adipocytes to the capillary endothelial cell surface. Glycosylphosphatidylinositol anchored high density lipoprotein binding protein 1 (GPIHBP1) binds LPL and transports it to the capillary lumen and anchors LPL to the capillary endothelium. Activation of LPL by Apo C-II, carried on the chylomicrons, leads to the hydrolysis of the triglycerides that are carried in the chylomicrons resulting in the formation of free fatty acids, which can be taken up by the adjacent muscle cells and adipocytes for either energy production or storage.  Fatty acid transport proteins (FATPs) and CD36 facilitate the uptake of fatty acids into adipocytes and muscle cells. Some of the free fatty acids released from chylomicrons bind to albumin and can be transported to other tissues. Apo A-V also plays an important role in activating LPL activity. Loss of function mutations in LPL, Apo C-II, GPIHPB1, lipase maturation factor 1, and Apo A-V can result in marked hypertriglyceridemia (familial chylomicronemia syndrome) (30). In addition, there are proteins that inhibit LPL activity. Apo C-III inhibits LPL activity and loss of function mutations in this gene are associated with increases in LPL activity and decreases in plasma triglyceride levels. Similarly, angiopoietin like protein 3 and 4, which target LPL for inactivation, regulate LPL activity. Loss of function mutations in these proteins also are associated with decreases in plasma triglyceride levels. Finally, the expression of LPL by muscle cells and adipocytes is regulated by hormones (particularly insulin), nutritional status, and inflammation.

 

The metabolism of the triglycerides carried in the chylomicrons results in a marked decrease in the size of these particles leading to the formation of chylomicron remnants, which are enriched in cholesterol esters and acquire Apo E. As these particles decrease in size phospholipids and apolipoproteins (Apo A and C) on the surface of the chylomicrons are transferred to other lipoproteins, mainly HDL. The transfer of Apo C-II from chylomicrons to HDL decreases the ability of LPL to further breakdown triglycerides. These chylomicron remnants are cleared from the circulation by the liver. The Apo E on the chylomicron remnants binds to the LDL receptor and other hepatic receptors such as LRP and syndecan-4 and the entire particle is taken up by the hepatocytes. Apo E is crucial for this process and mutations in Apo E (for example homozygosity for the Apo E2 isoform) can result in decreased chylomicron clearance and elevations in plasma cholesterol and triglyceride levels (familial dysbetalipoproteinemia) (30).

 

The exogenous lipoprotein pathway results in the efficient transfer of dietary fatty acids to muscle and adipose tissue for energy utilization and storage. The cholesterol is delivered to the liver where it can be utilized for the formation of VLDL, bile acids, or secreted back to the intestine via secretion into the bile. In normal individuals, this pathway can handle large amounts of fat (100 grams or more per day) without resulting in marked increases in plasma triglyceride levels. In fact, in a normal individual, a meal containing 75 grams of fat results in only a very modest increase in postprandial triglyceride levels. 

 

ENDOGENOUS LIPOPROTEIN PATHWAY (VLDL AND LDL)

 

Figure 7. Endogenous Lipoprotein Pathway

 

Formation of VLDL (50,63,64

 

In the liver triglycerides and cholesterol esters are transferred in the endoplasmic reticulum to newly synthesized Apo B-100. Similar to the intestine this transfer is mediated by MTTP. The availability of triglycerides is the primary determinant of the rate of VLDL synthesis. If the supply of triglyceride is limited the newly synthesized Apo B is rapidly degraded. Thus, in contrast to many proteins the rate of synthesis of the Apo B-100 is not the major determinant of the rate of secretion. Rather the amount of lipid available determines whether Apo B-100 is degraded or secreted. MTTP is required for the early addition of lipid to Apo B-100 particles but additional lipid is added via pathways that do not require MTTP. Additionally, the size of the VLDL particles is determined by the availability of triglycerides. When triglycerides are abundant the VLDL particles are large.

 

The quantity of fatty acids available for the synthesis of triglycerides is the main determinant of   triglyceride synthesis in the liver. The major sources of fatty acids are a) de novo fatty acid synthesis, b) the hepatic uptake of triglyceride rich lipoproteins, and c) the flux of fatty acids from adipose tissue to the liver. Diabetes, obesity, and the metabolic syndrome are common causes of an increase in hepatic triglyceride levels and the increased secretion of VLDL (44,65).

 

Loss of function mutations in either Apo B-100 or MTTP result in the failure to produce VLDL and marked decreases in plasma triglyceride and cholesterol levels (familial hypobetalipoproteinemia or abetalipoproteinemia) (56). The precise pathway by which the newly synthesized VLDL particles are secreted from the hepatocyte into the circulation is not resolved.

 

VLDL Metabolism (6,58)

 

VLDL particles are transported to peripheral tissues where the triglycerides are hydrolyzed by LPL and fatty acids are released. This process is very similar to that described above for chylomicrons and there is competition between the metabolism of chylomicrons and VLDL. High levels of chylomicrons can inhibit the clearance of VLDL. The removal of triglycerides from VLDL results in the formation of VLDL remnants (Intermediate density lipoproteins (IDL)). These IDL particles are relatively enriched in cholesterol esters and acquire Apo E from HDL particles. In a pathway analogous to the removal of chylomicron remnants these IDL particles can be removed from the circulation by the liver via binding of Apo E to LDL and LRP receptors. However, while the vast majority of chylomicron remnants are rapidly cleared from the circulation by the liver, only a fraction of IDL particles are cleared (approximately 50% but varies). The remaining triglycerides in the IDL particles are hydrolyzed by hepatic lipase leading to a further decrease in triglyceride content and the exchangeable apolipoproteins are transferred from the IDL particles to other lipoproteins leading to the formation of LDL. These LDL particles predominantly contain cholesterol esters and Apo B-100.  Thus, LDL is a product of VLDL metabolism.

 

LDL Metabolism (34,66-69)

 

The levels of plasma LDL are determined by the rate of LDL production and the rate of LDL clearance, both of which are regulated by the number of LDL receptors in the liver. The production rate of LDL from VLDL is partially determined by hepatic LDL receptor activity with a high LDL receptor activity resulting in a decrease in LDL production due to an increase in IDL uptake. Conversely, low LDL receptor activity results in an increase in LDL production formation due to a decrease in IDL uptake. With regards to LDL clearance, approximately 70% of circulating LDL is cleared via hepatocyte LDL receptor mediated endocytosis with the remainder taken up by extrahepatic tissues. An increase in the number of hepatic LDL receptors therefore increases LDL clearance leading to a decrease in plasma LDL levels. Conversely, a decrease in hepatic LDL receptors slows LDL clearance leading to an increase in plasma LDL levels. Thus, the level of hepatic LDL receptors plays a key role in regulating plasma LDL levels. Many of the drugs used to lower plasma LDL levels, such as the statins, ezetimibe, PCSK9 inhibitors, bile acid sequestrants and bempedoic acid lower plasma LDL levels by increasing the number of hepatic LDL receptors (57).

 

The levels of LDL receptors in the liver are mainly regulated by the cholesterol content of the hepatocyte. As cholesterol levels in the cell decrease, inactive sterol regulatory element binding proteins (SREBPs), which are transcription factors that mediate the expression of LDL receptors and key genes involved in cholesterol and fatty acid metabolism, are transported from the endoplasmic reticulum to the Golgi where proteases cleave the SREBPs into active transcription factors (Figure 4). These active SREBPs move to the nucleus where they stimulate the transcription of the LDL receptor and enzymes required for cholesterol synthesis, including HMG-CoA reductase, the rate limiting enzyme in cholesterol synthesis. If cholesterol levels in the cell are high, then the SREBPs remain in the endoplasmic reticulum in an inactive form and do not stimulate LDL receptor synthesis. In addition, cholesterol in the cell is oxidized and oxidized sterols activate LXR, a nuclear hormone receptor that is a transcription factor, which stimulates the transcription of E3 ubiquitin ligase that mediates the ubiquitination and degradation of the low-density lipoprotein receptor (Inducible degrader of the low-density lipoprotein receptor (IDOL)). Thus, the cell can sense the availability of cholesterol and regulate LDL receptor activity. If the cholesterol content of the cell is decreased LDL receptor activity is increased to allow for the increased uptake of cholesterol. Conversely, if the cholesterol content of the cell is increased LDL receptor activity is decreased and the uptake of LDL by the cell is diminished. Statins, ezetimibe, bile acid sequestrants, and bempedoic acid decrease hepatic cholesterol levels thereby increasing LDL receptor levels and decreasing plasma LDL levels (57). Finally, the LDL receptor is targeted for degradation by PCSK9, a secreted protein that binds to the LDL receptor and enhances LDL receptor degradation in the lysosomes. Loss of function mutations in PCSK9 and drugs that inhibit PCSK9 result in increased LDL receptor activity and decreased LDL levels while gain of function mutations in PCSK9 lead to decreased LDL receptor activity and elevations in LDL levels.

 

Thus, the endogenous lipoprotein pathway facilitates the movement of triglycerides synthesized in the liver to muscle and adipose tissue. Additionally, it also provides a pathway for the transport of cholesterol from the liver to peripheral tissues.

 

HDL METABOLISM AND REVERSE CHOLESTEROL TRANSPORT (38-40,47,48,70,71)

 

Figure 8. HDL Metabolism

 

HDL Formation

 

Several steps are required to generate mature HDL particles. The first step involves the synthesis of the main structural protein contained in HDL, Apo A-I. Apo A-I is synthesized predominantly by the liver and intestine. After Apo A-I is secreted, it acquires cholesterol and phospholipids that are effluxed from hepatocytes and enterocytes. The efflux of cholesterol and phospholipids to the newly synthesized lipid poor Apo A-I (pre-beta HDL) is facilitated by ABCA1. Patients with loss of function mutations in ABCA1 (Tangiers disease) fail to lipidate the newly secreted Apo A-I leading to the rapid catabolism of Apo A-I and very low HDL levels (72). Using mice with targeted knock-out of ABCA1 it has been shown that HDL cholesterol levels are reduced by 80% in mice lacking ABCA1 in the liver and 30% in mice lacking ABCA1 in the intestine. While initially cholesterol and phospholipids are obtained from the liver and intestine, HDL also acquires lipid from other tissues and from other lipoproteins. Muscle cells, adipocytes, and other tissues express ABCA1 and ABCG1 and are able to transfer cholesterol and phospholipids to Apo A-I particles. Additionally, as noted above, newly formed HDL can also obtain cholesterol and phospholipids from chylomicrons and VLDL during their lipolysis by LPL. This accounts for the observation that patients with high plasma triglyceride levels due to decreased clearance frequently have low HDL cholesterol levels. Additionally, phospholipid transfer protein (PLTP) facilitates the movement of phospholipids between lipoproteins; mice lacking PLTP have a marked reduction in HDL cholesterol and Apo A-I levels. Finally, the lipolysis of triglyceride rich lipoproteins also results in the transfer of apolipoproteins from these particles to HDL.

 

HDL Cholesterol Esterification

 

As noted earlier the cholesterol in the core of HDL is esterified (cholesterol esters). The cholesterol that is effluxed from cells to HDL is free cholesterol and is localized on the surface of HDL particles. In order to form mature large spherical HDL particles with a core of cholesterol esters the free cholesterol transferred from cells to the surface of HDL particles must be esterified. LCAT, an HDL associated enzyme catalyzes the transfer of a fatty acid from phospholipids to free cholesterol resulting in the formation of cholesterol esters. The cholesterol ester formed is then able to move from the surface of the HDL particle to the core allowing additional free cholesterol to be transferred from cells to HDL particles. Apo A-I is an activator of LCAT and facilitates this esterification process. LCAT activity is required for the formation of large HDL particles. LCAT deficiency in humans results in decreased HDL cholesterol and Apo A-I levels and a higher percentage of small HDL particles (72).

 

HDL Metabolism

 

Lipases and transfer proteins play an important role in determining the size and composition of HDL particles. The cholesterol ester carried in the core of HDL particles may be transferred to Apo B containing particles in exchange for triglyceride. This transfer is mediated by CETP and results in HDL enriched in triglycerides which may then be metabolized by lipases. Humans deficient in CETP activity have very high HDL cholesterol levels and large HDL particles (72). CETP also impacts LDL cholesterol levels and the absence of CETP results in a decrease in LDL cholesterol. Mice do not have CETP and have relatively high HDL cholesterol levels and low LDL cholesterol levels. Hepatic lipase hydrolyzes both triglycerides and phospholipids in HDL. The triglycerides that are transferred to HDL by CETP activity are catabolized by hepatic lipase resulting in the formation of small HDL particles and Apo A-I more easily disassociates from small HDL resulting in the release of Apo A-I and increased Apo A-I degradation. Genetic deficiency of hepatic lipase results in a modest elevation in HDL cholesterol levels and larger HDL particles (72). Hepatic lipase activity is increased in insulin resistant states and this is associated with reduced HDL cholesterol levels. Endothelial cell lipase is a phospholipase that hydrolyzes the phospholipids carried in HDL particles. In mice increased endothelial lipase activity results in decreased HDL cholesterol levels while decreased endothelial lipase activity increases HDL cholesterol levels.

 

The cholesterol carried on HDL is primarily delivered to the liver. The uptake of HDL cholesterol by the liver is mediated by SR-BI, which promotes the selective uptake of HDL cholesterol. The HDL particle binds to SR-BI and the cholesterol in HDL is transported into the liver without internalization of the HDL particle. A smaller cholesterol depleted HDL particle is formed, which is then released back into the circulation. In SR-BI deficient mice there is a marked increase in HDL cholesterol levels. Interestingly the risk of atherosclerosis is increased in these SR-BI deficient mice despite an increase in HDL cholesterol levels. Notably, while HDL cholesterol levels are increased in SR-B1 deficient mice the reverse cholesterol transport pathway is actually reduced. While in mice the physiological importance of the hepatic SR-BI pathway is clear, the role in humans is uncertain. In mice, the movement of cholesterol from peripheral tissues to the liver is dependent solely on SR-BI while in humans CETP can facilitate the transport of cholesterol from HDL to Apo B containing lipoproteins, which serves as an alternative pathway for the transport cholesterol to the liver. 

 

Apo A-I is metabolized independently of HDL cholesterol. Most of the Apo A-I is catabolized by the kidneys with the remainder catabolized by the liver. Lipid free or lipid poor Apo A-I is filtered by the kidneys and then taken up by the renal tubules. The size of the Apo A-I particle determines whether it can be filtered by the kidneys and hence the degree of lipidation of Apo A-I determines the rate of catabolism. Conditions or disease states (for example Tangiers disease, which is due to a mutation in ABCA1, or LCAT deficiency) that result in lipid poor HDL led to the accelerated catabolism of Apo A-I by the kidney. Apo A-I binds to cubilin, which in conjunction with megalin, a member of the LDL receptor gene family, leads to the uptake and degradation of filtered Apo A-I by renal tubular cells. While the liver is also involved in the catabolism of Apo A-I, the mechanisms are poorly understood. HDL particles may contain Apo E and it is therefore possible that Apo E containing HDL particles are taken up via the LDL receptor and other Apo E receptors in the liver and degraded.

 

Reverse Cholesterol Transport (73-78)

 

Peripheral cells accumulate cholesterol through the uptake of circulating lipoproteins and de novo cholesterol synthesis. Most cells do not have a mechanism for catabolizing cholesterol. Cells that synthesize steroid hormones can convert cholesterol to glucocorticoids, estrogen, testosterone, etc. Intestinal cells, sebocytes, and keratinocytes can secrete cholesterol into the intestinal lumen or onto the skin surface thereby eliminating cholesterol. However, in order for most cells to decrease their cholesterol content reverse cholesterol transport is required. From a clinical point of view, the ability of macrophages in the arterial wall to efficiently efflux cholesterol into the reverse cholesterol transport pathway may play an important role in the prevention of atherosclerosis.

 

As noted earlier ABCA1 plays an important role in the efflux of cholesterol to lipid poor pre-beta Apo A-I particles (Figure 9). ABCG1 plays an important role in the efflux of cholesterol from cells to mature HDL particles. In some studies, SR-B1 also plays a role in the efflux of cholesterol to mature HDL particles. Additionally, passive diffusion of cholesterol from the plasma membrane to HDL may also contribute to cholesterol efflux. The levels of both ABCA1 and ABCG1 are increased by LXR activation. LXR is a nuclear hormone transcription factor that is activated by oxysterols. As the cholesterol levels in a cell increase the formation of oxysterols increases leading to the activation of LXR resulting in an increase in ABCA1 and ABCG1 expression, which will result in the enhanced efflux of cholesterol from the cell to HDL.  Additionally, ABCA1 and ABCG1 mRNAs are targeted for degradation by miR-33, a microRNA that is embedded within the SREBP2 gene. An increase in cellular cholesterol decreases the expression of SREBP2 leading to a decrease in miR-33 resulting in enhanced LXR expression. Thus, the decrease in SREBP2 transcription will lead to a decrease in LDL receptor activity and a reduction in cholesterol uptake, while simultaneously, a decrease in miR-33 will lead to an increase in LXR activity stimulating the expression of ABCA1 and ABCG1 resulting in increased cholesterol efflux. Conversely a decrease in cellular cholesterol levels will increase SREBP2 expression resulting in an increase in LDL receptor activity and an increase in miR-33, which will result in a decrease in LXR activity, decreased expression of ABCA1 and ABCG1, and a reduction in cholesterol efflux. Together changes in cholesterol uptake mediated by the LDL receptor and cholesterol efflux mediated by ABCA1 and ABCG1 will maintain cellular cholesterol homeostasis.

 

Figure 9. Cholesterol Efflux from Macrophages (modified from J. Clinical Investigation 116: 3090, 2006)

 

Once cholesterol is transferred from cells to HDL there are two pathways for the cholesterol to be transported and taken up by the liver. As discussed earlier, HDL can interact with hepatic SR-BI receptors resulting in the selective uptake of cholesterol from HDL particles. Alternatively, CETP can transfer cholesterol from HDL particles to Apo B containing particles with the subsequent uptake of Apo B containing lipoproteins by the liver. After the delivery of cholesterol to the liver there are several pathways by which the cholesterol can be eliminated. Cholesterol can be converted to bile acids and secreted in the bile. Alternatively, cholesterol can be directly secreted into the bile. ABCG5 and ABCG8 promote the transport of cholesterol into the bile and the expression of these genes is enhanced by LXR activation. Thus, an increase in hepatic cholesterol levels leading to increased oxysterol production will activate LXR resulting in the increased expression of ABCG5 and ABCG8 facilitating the secretion of cholesterol in the bile.

 

Evidence suggests that reverse cholesterol transport plays an important role in protecting from the development of atherosclerosis. It should be noted that HDL cholesterol levels may not be indicative of the rate of reverse cholesterol transport. As described above reverse cholesterol transport involves several steps and the level of HDL cholesterol may not accurately reflect these steps. For example, studies have shown that the ability of HDL to promote cholesterol efflux from macrophages can vary. Thus, the same level of HDL cholesterol may not have equivalent abilities to mediate the initial step of reverse cholesterol transport.  

 

LIPOPROTEIN (a) (14-16,79)

 

Figure 10. Lp (a)

 

Lp (a) consists of an LDL molecule and a unique apolipoprotein (a), which is attached to the Apo B-100 of the LDL via a single disulfide bound. Lp (a) contain Apo (a) and Apo B-100 in a 1:1 molar ratio. Like Apo B-100, apo (a) is also made by hepatocytes. Apo (a) contains multiple kringle motifs that are similar to the kringle repeats in plasminogen. The number of kringle repeats can vary and thus the molecular weight of apo (a) can range from 250,000 to 800,000.  The levels of Lp (a) in plasma can vary more than a 1000-fold ranging from undetectable to greater than 100mg/dl. Lp (a) levels largely reflect Lp (a) production rates, which are primarily genetically regulated and not greatly affected by environmental factors. Individuals with high molecular weight Apo (a) proteins tend to have lower levels of Lp (a) while individuals with low molecular weight Apo (a) tend to have higher levels. It is hypothesized that the liver is less efficient in secreting high molecular weight Apo (a). The mechanism of Lp (a) clearance is uncertain but does not appear to primarily involve LDL receptors. Therapies that accelerate LDL clearance and lower LDL levels do not lower Lp (a) levels (for example statin therapy). The kidney appears to play an important role in Lp (a) clearance as kidney disease is associated with delayed clearance and elevations in Lp (a) levels.

 

 Elevated plasma Lp(a) levels are associated with an increased risk of atherosclerosis. Apo (a) is an inhibitor of fibrinolysis and enhances the uptake of lipoproteins by macrophages, both of which could account for the increased the risk of atherosclerosis in individuals with elevated Apo (a) levels. Additionally, Lp (a) is the major lipoprotein carrier of oxidized phospholipids, which are inflammatory and could also increase the risk of atherosclerosis. The physiologic function of Apo (a) is unknown. Apo (a) is found in primates but not in other species.

 

ACKNOWLEDGEMENTS

 

This work was supported by grants from the Northern California Institute for Research and Education.

 

REFERENCES

 

  1. Feingold KR, Grunfeld C. Lipids: a key player in the battle between the host and microorganisms. J Lipid Res 2012; 53:2487-2489
  2. Nielsen LB, Nielsen MJ, Moestrup SK. Lipid metabolism: an apolipoprotein-derived weapon combating trypanosoma infection. Curr Opin Lipidol 2006; 17:699-701
  3. Feingold KR. The bidirectional link between HDL and COVID-19 infections. J Lipid Res 2021; 62:100067
  4. Smith LC, Pownall HJ, Gotto AM, Jr. The plasma lipoproteins: structure and metabolism. Annu Rev Biochem 1978; 47:751-757
  5. Julve J, Martin-Campos JM, Escola-Gil JC, Blanco-Vaca F. Chylomicrons: Advances in biology, pathology, laboratory testing, and therapeutics. Clin Chim Acta 2016; 455:134-148
  6. Chait A, Ginsberg HN, Vaisar T, Heinecke JW, Goldberg IJ, Bornfeldt KE. Remnants of the Triglyceride-Rich Lipoproteins, Diabetes, and Cardiovascular Disease. Diabetes 2020; 69:508-516
  7. Krauss RM, King SM. Remnant lipoprotein particles and cardiovascular disease risk. Best Pract Res Clin Endocrinol Metab 2023; 37:101682
  8. Berneis KK, Krauss RM. Metabolic origins and clinical significance of LDL heterogeneity. J Lipid Res 2002; 43:1363-1379
  9. Asztalos BF, Niisuke K, Horvath KV. High-density lipoprotein: our elusive friend. Curr Opin Lipidol 2019; 30:314-319
  10. Thakkar H, Vincent V, Sen A, Singh A, Roy A. Changing Perspectives on HDL: From Simple Quantity Measurements to Functional Quality Assessment. J Lipids 2021; 2021:5585521
  11. Thomas SR, Zhang Y, Rye KA. The pleiotropic effects of high-density lipoproteins and apolipoprotein A-I. Best Pract Res Clin Endocrinol Metab 2023; 37:101689
  12. Vaisar T, Pennathur S, Green PS, Gharib SA, Hoofnagle AN, Cheung MC, Byun J, Vuletic S, Kassim S, Singh P, Chea H, Knopp RH, Brunzell J, Geary R, Chait A, Zhao XQ, Elkon K, Marcovina S, Ridker P, Oram JF, Heinecke JW. Shotgun proteomics implicates protease inhibition and complement activation in the antiinflammatory properties of HDL. J Clin Invest 2007; 117:746-756
  13. Kostner KM, Kostner GM. Lipoprotein (a): a historical appraisal. J Lipid Res 2017; 58:1-14
  14. Nordestgaard BG, Langsted A. Lipoprotein (a) as a cause of cardiovascular disease: insights from epidemiology, genetics, and biology. J Lipid Res 2016; 57:1953-1975
  15. Schmidt K, Noureen A, Kronenberg F, Utermann G. Structure, function, and genetics of lipoprotein (a). J Lipid Res 2016; 57:1339-1359
  16. Khovidhunkit W. Lipoprotein(a). In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  17. Mahley RW, Innerarity TL, Rall SC, Jr., Weisgraber KH. Plasma lipoproteins: apolipoprotein structure and function. J Lipid Res 1984; 25:1277-1294
  18. Breslow JL. Human apolipoprotein molecular biology and genetic variation. Annu Rev Biochem 1985; 54:699-727
  19. Frank PG, Marcel YL. Apolipoprotein A-I: structure-function relationships. J Lipid Res 2000; 41:853-872
  20. Chan DC, Ng TW, Watts GF. Apolipoprotein A-II: evaluating its significance in dyslipidaemia, insulin resistance, and atherosclerosis. Ann Med 2012; 44:313-324
  21. Wang F, Kohan AB, Lo CM, Liu M, Howles P, Tso P. Apolipoprotein A-IV: a protein intimately involved in metabolism. J Lipid Res 2015; 56:1403-1418
  22. Hubacek JA. Apolipoprotein A5 fifteen years anniversary: Lessons from genetic epidemiology. Gene 2016; 592:193-199
  23. Sharma V, Forte TM, Ryan RO. Influence of apolipoprotein A-V on the metabolic fate of triacylglycerol. Curr Opin Lipidol 2013; 24:153-159
  24. Anant S, Davidson NO. Molecular mechanisms of apolipoprotein B mRNA editing. Curr Opin Lipidol 2001; 12:159-165
  25. Levenson AE, de Ferranti SD. Familial Hypercholesterolemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  26. Wolska A, Dunbar RL, Freeman LA, Ueda M, Amar MJ, Sviridov DO, Remaley AT. Apolipoprotein C-II: New findings related to genetics, biochemistry, and role in triglyceride metabolism. Atherosclerosis 2017; 267:49-60
  27. Ramms B, Gordts P. Apolipoprotein C-III in triglyceride-rich lipoprotein metabolism. Curr Opin Lipidol 2018; 29:171-179
  28. D'Erasmo L, Di Costanzo A, Gallo A, Bruckert E, Arca M. ApoCIII: A multifaceted protein in cardiometabolic disease. Metabolism 2020; 113:154395
  29. Wolska A, Reimund M, Remaley AT. Apolipoprotein C-II: the re-emergence of a forgotten factor. Curr Opin Lipidol 2020; 31:147-153
  30. Patni N, Ahmad Z, Wilson DP. Genetics and Dyslipidemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  31. Taskinen MR, Boren J. Why Is Apolipoprotein CIII Emerging as a Novel Therapeutic Target to Reduce the Burden of Cardiovascular Disease? Curr Atheroscler Rep 2016; 18:59
  32. Witztum JL, Gaudet D, Freedman SD, Alexander VJ, Digenio A, Williams KR, Yang Q, Hughes SG, Geary RS, Arca M, Stroes ESG, Bergeron J, Soran H, Civeira F, Hemphill L, Tsimikas S, Blom DJ, O'Dea L, Bruckert E. Volanesorsen and Triglyceride Levels in Familial Chylomicronemia Syndrome. N Engl J Med 2019; 381:531-542
  33. Mahley RW. Apolipoprotein E: from cardiovascular disease to neurodegenerative disorders. J Mol Med (Berl) 2016; 94:739-746
  34. Goldstein JL, Brown MS. The LDL receptor. Arterioscler Thromb Vasc Biol 2009; 29:431-438
  35. Goldstein JL, DeBose-Boyd RA, Brown MS. Protein sensors for membrane sterols. Cell 2006; 124:35-46
  36. van de Sluis B, Wijers M, Herz J. News on the molecular regulation and function of hepatic low-density lipoprotein receptor and LDLR-related protein 1. Curr Opin Lipidol 2017; 28:241-247
  37. Krauss RM, Lu JT, Higgins JJ, Clary CM, Tabibiazar R. VLDL receptor gene therapy for reducing atherogenic lipoproteins. Mol Metab 2023; 69:101685
  38. Trigatti BL. SR-B1 and PDZK1: partners in HDL regulation. Curr Opin Lipidol 2017; 28:201-208
  39. Wang S, Smith JD. ABCA1 and nascent HDL biogenesis. Biofactors 2014; 40:547-554
  40. Baldan A, Tarr P, Lee R, Edwards PA. ATP-binding cassette transporter G1 and lipid homeostasis. Curr Opin Lipidol 2006; 17:227-232
  41. Kidambi S, Patel SB. Cholesterol and non-cholesterol sterol transporters: ABCG5, ABCG8 and NPC1L1: a review. Xenobiotica 2008; 38:1119-1139
  42. Patel SB, Graf GA, Temel RE. ABCG5 and ABCG8: more than a defense against xenosterols. J Lipid Res 2018; 59:1103-1113
  43. Olivecrona G. Role of lipoprotein lipase in lipid metabolism. Curr Opin Lipidol 2016; 27:233-241
  44. Feingold KR. Dyslipidemia in Patients with Diabetes. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  45. Kobayashi J, Miyashita K, Nakajima K, Mabuchi H. Hepatic Lipase: a Comprehensive View of its Role on Plasma Lipid and Lipoprotein Metabolism. J Atheroscler Thromb 2015; 22:1001-1011
  46. Yasuda T, Ishida T, Rader DJ. Update on the role of endothelial lipase in high-density lipoprotein metabolism, reverse cholesterol transport, and atherosclerosis. Circ J 2010; 74:2263-2270
  47. Ossoli A, Simonelli S, Vitali C, Franceschini G, Calabresi L. Role of LCAT in Atherosclerosis. J Atheroscler Thromb 2016; 23:119-127
  48. Mabuchi H, Nohara A, Inazu A. Cholesteryl ester transfer protein (CETP) deficiency and CETP inhibitors. Mol Cells 2014; 37:777-784
  49. Shrestha S, Wu BJ, Guiney L, Barter PJ, Rye KA. Cholesteryl ester transfer protein and its inhibitors. J Lipid Res 2018; 59:772-783
  50. Hooper AJ, Burnett JR, Watts GF. Contemporary aspects of the biology and therapeutic regulation of the microsomal triglyceride transfer protein. Circ Res 2015; 116:193-205
  51. Abumrad NA, Davidson NO. Role of the gut in lipid homeostasis. Physiol Rev 2012; 92:1061-1085
  52. D'Aquila T, Hung YH, Carreiro A, Buhman KK. Recent discoveries on absorption of dietary fat: Presence, synthesis, and metabolism of cytoplasmic lipid droplets within enterocytes. Biochim Biophys Acta 2016; 1861:730-747
  53. Hussain MM. Intestinal lipid absorption and lipoprotein formation. Curr Opin Lipidol 2014; 25:200-206
  54. Kindel T, Lee DM, Tso P. The mechanism of the formation and secretion of chylomicrons. Atheroscler Suppl 2010; 11:11-16
  55. Liebeskind A, Peterson AL, Wilson D. Sitosterolemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  56. Shapiro MD, Feingold KR. Monogenic Disorders Causing Hypobetalipoproteinemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  57. Feingold KR. Cholesterol Lowering Drugs. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  58. Dallinga-Thie GM, Franssen R, Mooij HL, Visser ME, Hassing HC, Peelman F, Kastelein JJ, Peterfy M, Nieuwdorp M. The metabolism of triglyceride-rich lipoproteins revisited: new players, new insight. Atherosclerosis 2010; 211:1-8
  59. Dijk W, Kersten S. Regulation of lipid metabolism by angiopoietin-like proteins. Curr Opin Lipidol 2016; 27:249-256
  60. Fong LG, Young SG, Beigneux AP, Bensadoun A, Oberer M, Jiang H, Ploug M. GPIHBP1 and Plasma Triglyceride Metabolism. Trends Endocrinol Metab 2016; 27:455-469
  61. Peterfy M. Lipase maturation factor 1: a lipase chaperone involved in lipid metabolism. Biochim Biophys Acta 2012; 1821:790-794
  62. Young SG, Fong LG, Beigneux AP, Allan CM, He C, Jiang H, Nakajima K, Meiyappan M, Birrane G, Ploug M. GPIHBP1 and Lipoprotein Lipase, Partners in Plasma Triglyceride Metabolism. Cell Metab 2019; 30:51-65
  63. Tiwari S, Siddiqi SA. Intracellular trafficking and secretion of VLDL. Arterioscler Thromb Vasc Biol 2012; 32:1079-1086
  64. Choi SH, Ginsberg HN. Increased very low density lipoprotein (VLDL) secretion, hepatic steatosis, and insulin resistance. Trends Endocrinol Metab 2011; 22:353-363
  65. Feingold KR. Obesity and Dyslipidemia. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  66. Goldstein JL, Brown MS. A century of cholesterol and coronaries: from plaques to genes to statins. Cell 2015; 161:161-172
  67. Horton JD, Cohen JC, Hobbs HH. Molecular biology of PCSK9: its role in LDL metabolism. Trends Biochem Sci 2007; 32:71-77
  68. Zhang L, Reue K, Fong LG, Young SG, Tontonoz P. Feedback regulation of cholesterol uptake by the LXR-IDOL-LDLR axis. Arterioscler Thromb Vasc Biol 2012; 32:2541-2546
  69. Brown MS, Radhakrishnan A, Goldstein JL. Retrospective on Cholesterol Homeostasis: The Central Role of Scap. Annual review of biochemistry 2017;
  70. Rosenson RS, Brewer HB, Jr., Davidson WS, Fayad ZA, Fuster V, Goldstein J, Hellerstein M, Jiang XC, Phillips MC, Rader DJ, Remaley AT, Rothblat GH, Tall AR, Yvan-Charvet L. Cholesterol efflux and atheroprotection: advancing the concept of reverse cholesterol transport. Circulation 2012; 125:1905-1919
  71. Rye KA, Barter PJ. Cardioprotective functions of HDLs. J Lipid Res 2014; 55:168-179
  72. Shapiro MD, Feingold KR. Monogenic Disorders Altering HDL Levels. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, de Herder WW, Dhatariya K, Dungan K, Hofland J, Kalra S, Kaltsas G, Kapoor N, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, Levy M, McGee EA, McLachlan R, New M, Purnell J, Sahay R, Shah AS, Singer F, Sperling MA, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  73. Zhao Y, Van Berkel TJ, Van Eck M. Relative roles of various efflux pathways in net cholesterol efflux from macrophage foam cells in atherosclerotic lesions. Curr Opin Lipidol 2010; 21:441-453
  74. Lee-Rueckert M, Escola-Gil JC, Kovanen PT. HDL functionality in reverse cholesterol transport--Challenges in translating data emerging from mouse models to human disease. Biochim Biophys Acta 2016; 1861:566-583
  75. Tall AR. Cholesterol efflux pathways and other potential mechanisms involved in the athero-protective effect of high density lipoproteins. J Intern Med 2008; 263:256-273
  76. Siddiqi HK, Kiss D, Rader D. HDL-cholesterol and cardiovascular disease: rethinking our approach. Curr Opin Cardiol 2015; 30:536-542
  77. Moore KJ, Rayner KJ, Suarez Y, Fernandez-Hernando C. The role of microRNAs in cholesterol efflux and hepatic lipid metabolism. Annu Rev Nutr 2011; 31:49-63
  78. Ouimet M, Barrett TJ, Fisher EA. HDL and Reverse Cholesterol Transport. Circ Res 2019; 124:1505-1518
  79. Hoover-Plow J, Huang M. Lipoprotein(a) metabolism: potential sites for therapeutic targets. Metabolism 2013; 62:479-491

Monogenic Disorders Causing Hypobetalipoproteinemia

ABSTRACT

 

Monogenic mutations leading to hypobetalipoproteinemia are rare. The monogenic causes of hypobetalipoproteinemia include familial hypobetalipoproteinemia, abetalipoproteinemia, chylomicron retention disease, loss of function mutations in PCSK9, and loss of function mutations in angiopoietin-like protein 3 (ANGPTL3) (Familiar Combined Hypolipidemia). This chapter describes the etiology, pathogenesis, clinical and laboratory findings, and the treatment of these rare monogenic disorders.

 

INTRODUCTION

 

Monogenic mutations leading to hypobetalipoproteinemia are rare. The monogenic causes of hypobetalipoproteinemia include familial hypobetalipoproteinemia (FHBL), abetalipoproteinemia (ABL), chylomicron retention disease (CMRD), loss of function mutations in PCSK9, and loss of function mutations in angiopoietin-like protein 3 (ANGPTL3) (Familial Combined Hypolipidemia, FCH) (1). Increased understanding of the genetic and the molecular underpinnings of these disorders has allowed a focused prioritization of therapeutic targets for drug development. Table 1 summarizes genetic, lipid, and clinical features of the major hypobetalipoproteinemia syndromes and table 2 provides a new classification of these disorders. Of note the parental lipid profile is normal in abetalipoproteinemia and chylomicron retention disease.

 

It should be recognized that secondary, non-familial, forms of hypobetalipoproteinemia occur and include strict vegan diet, malnutrition, malabsorption, hyperthyroidism, malignancy, and chronic liver disease. In addition, hypobetalipoproteinemia can also be due to polymorphisms in multiple genes that together result in hypobetalipoproteinemia (polygenic etiology) (2-4). In a study of 111 patients with LDL-C levels below the fifth percentile 36% had monogenic hypobetalipoproteinemia, 34% had polygenic hypobetalipoproteinemia, and 30% had hypobetalipoproteinemia from an unknown cause (2). In a study of women with an LDL-C ≤1st percentile (≤50 mg/dL) 15.7% carried mutations causing monogenic hypocholesterolemia and 49.6% were genetically predisposed to a low LDL-C on the basis of an extremely low weighted polygenetic risk score (4). Of note individuals with monogenic hypobetalipoproteinemia are more likely to have liver steatosis than individuals without a monogenic disorder (2).

 

Table 1. Characteristics of the Hypobetalipoproteinemia Syndromes

 

Inheritance

Effected gene

Prevalence

Lipids

Clinical features

FHBL

ACD

Truncation mutations in Apo B

1:1000 – 1:3000

Apo B <5th percentile,

LDL-C 20- 50 mg/dL

Hepatic steatosis

Mild elevation of transaminases. Lower prevalence of ASCVD

ABL

 

FHBL

AR

 

AR

MTTP

 

Apo B

<1:1,000,000

Triglycerides < 30 mg/dL,

Cholesterol < 30 mg/dL),

LDL and Apo B undetectable

Hepatic steatosis

Malabsorption, steatorrhea, diarrhea, and failure to thrive.

Deficiency of fat-soluble vitamins.

PCSK9

ACD

Loss of function mutations in PCSK9

 

Heterozygous – mild to moderate reduction in LDL-C

Homozygous – LDL-C ~15 mg/dL

Normal health; significantly lower prevalence of ASCVD

FCH

ACD

Loss of function mutations in ANGPTL3

Very rare

Panhypolipidemia

Normal health; significantly lower prevalence of ASCVD

CMRD

AR

SAR1B

Very rare

LDL-C and HDL-C -decreased by 50%,

Triglycerides - normal

hypocholesterolemia associated with failure to thrive, diarrhea, steatorrhea, and abdominal distension

ACD- autosomal co-dominant; AR- autosomal recessive; FHBL- familial hypobetalipoproteinemia; ABL- abetalipoproteinemia; FCH- Familiar Combined Hypolipidemia; CMRD- chylomicron retention disease, MTTP- microsomal triglyceride transfer protein; ANGPTL3- angiopoietin-like protein 3; ASCVD- atherosclerotic cardiovascular disease.

 

Table 2. Classification of Disorders Causing Familial Hypocholesterolemia

New Name

Common Name

Gene Defect

Class I: Familial hypobetalipoproteinemia due to lipoprotein assembly and secretion defects

FHBL-SD1

Abetalipoproteinemia

Microsomal Triglyceride Transfer Protein

FHBL-SD2

Familial Hypobetalipoproteinemia

Apolipoprotein B

FHBL-SD3

Chylomicron retention disease

SAR1B

Class II: Familial hypobetalipoproteinemia due to enhanced lipoprotein catabolism

FHBL-EC1

Familial Combined Hypolipidemia

ANGPTL3

FHBL-EC2

 

PCSK9

Modified from (5).

 

FAMILIAL HYPOBETALIPOPROTEINEMIA  

 

Familial Hypobetalipoproteinemia (FHBL) is a relatively common autosomal semi-dominant disorder most commonly due to truncation mutations in the gene coding for Apo B (1,6-8). The prevalence of heterozygous FHBL is estimated to be 1 in 700 to 3000 (1). Variants that lead to truncated proteins that are 30% in length or shorter have more severe signs and symptoms than those with longer truncated proteins (6,7). The truncated forms of Apo B found in FHBL are generally non-functional (truncation decreases lipidation and secretion) and are catabolized quickly, resulting in markedly reduced levels in the plasma (Apo B <5th percentile and LDL-C typically between 20- 50 mg/dL) (7,8). Although there is one normal allele in heterozygous FHBL, plasma Apo B levels are approximately 25% of normal rather than the predicted 50% (8). These lower-than-expected levels result from a lower secretion rate of VLDL Apo B from the liver, decreased production of LDL Apo B, increased catabolism of VLDL, and extremely low secretion of the truncated Apo B (6-8). Given the reduced substrate (Apo B) for lipid (predominantly triglyceride) loading, fatty liver develops in these patients (6,9). Hepatic steatosis and mild elevation of liver enzymes are common in heterozygous FHBL (6,9). Interestingly, individuals with monogenic hypobetalipoproteinemia had a much greater prevalence of hepatic dysfunction than individuals with polygenic hypobetalipoproteinemia (2). In contrast to non-alcoholic fatty liver disease, FHBL is not associated with hepatic or peripheral insulin resistance (9). This observation, however, does not imply that hepatic steatosis associated with FHBL is benign. There are several reports of steatohepatitis, cirrhosis, and hepatocellular carcinoma in patients with FHBL and it is estimated that 5-10% of individuals with FHBL develop relatively more severe nonalcoholic steatohepatitis (6). Because of the risk of developing liver disease liver function tests should be checked every 1-2 years and a hepatic ultrasound in those with elevated liver transaminases (6). While hepatic fat accumulation is the rule, there is generally sufficient chylomicron production to handle dietary fat. However, oral fat intolerance and intestinal fat malabsorption have been reported (6). On the positive side the decrease in proatherogenic lipoproteins has been associated with a reduced risk of cardiovascular disease (10).

 

Given the association of FHBL and low LDL-C, Apo B has been an attractive target for drug development. Indeed, unraveling the genetic and molecular mechanisms of FHBL provided the motivation to pharmacologically antagonize Apo B synthesis for therapeutic gains. This culminated in the production of mipomersen, a synthetic single strand anti-sense oligonucleotide to Apo B (11,12). Essentially, anti-sense oligonucleotides contain approximately ~20 deoxyribonucleic acid (DNA) base pairs complementary to a unique messenger ribonucleic acid (mRNA) sequence. The hybridization of the anti-sense oligonucleotide to the mRNA of interest leads to its catabolism via RNase H1, with markedly reduced mRNA levels and ultimately reduced target protein levels. In this case, mipomersen binds to Apo B mRNA leading to reduced production of the protein, and mimicking (albeit to a lesser extent) FHBL. Mipomersen is the first anti-sense oligonucleotide approved by the United States Food and Drug Administration (FDA) and was commercialized in 2013 with a limited indication for adjunctive LDL-C lowering in patients with homozygous familial hypercholesterolemia (HoFH) (12). It is an injectable agent administered subcutaneously once a week. In the clinical trials, mipomersen was associated with a reduction of LDL-C by 21% in subjects with HoFH and 33% in subjects with heterozygous familial hypercholesterolemia (HeFH) (12). Interestingly, it was also found to lower Lp(a) by 21- 23% (12). While it is highly efficacious in LDL-C lowering, it has side effects, many of which can be predicted based on the experience with FHBL (e.g., hepatic steatosis, elevated liver enzymes) (12). It is also associated with injection site reactions in a considerable number of subjects (12). In May 2018 sales were discontinued due to safety concerns related to increased liver transaminases and fatty liver.

 

Homozygous hypobetalipoproteinemia (HHBL) is extremely rare (6). These patients are homozygous or compound heterozygous for mutations in the Apo B gene. The clinical manifestations mimic ABL (see below) (6).

 

ABETALIPOPROTEINEMIA  

 

Abetalipoproteinemia (ABL) is a rare autosomal recessive disorder characterized by very low plasma concentrations of triglyceride and cholesterol (under 30 mg/dL) and undetectable levels of LDL and Apo B (1,7,13,14). The incidence of ABL is < 1 in 1,000,000. HDL-C levels are usually normal or modestly reduced. It is due to mutations in the gene that codes for microsomal triglyceride transfer protein (MTTP) (7,13-15). MTTP lipidates nascent Apo B in the endoplasmic reticulum to produce VLDL and chylomicrons in the liver and small intestine, respectively (15,16). Unlipidated Apo B is targeted for proteasomal degradation leading to the absence of Apo B containing lipoproteins in the plasma (and thus markedly reduced levels of LDL-C and triglycerides) (15,16). Similar to FHBL, VLDL production is inhibited (14). The reduced triglyceride export from the liver leads to hepatic steatosis, which rarely may progress to steatohepatitis, fibrosis, and cirrhosis (1,9,13). Additionally, lack of MTTP facilitated lipidation of chylomicrons in the small intestine results in lipid accumulation in enterocytes with associated malabsorption, steatorrhea, and diarrhea (1,7,13). The malabsorption and diarrhea lead to failure to thrive during infancy (1,7,13). A decrease in dietary fat can reduce the gastrointestinal symptoms. Acanthocytosis may encompass 50% of circulating red blood cells (red blood cells with spiked cell membranes, due to thorny projections) due to alterations in the lipid composition and fluidity of red cell membranes (1,13,14). An additional issue of importance related to ABL is deficiency of fat-soluble vitamins (1,13). Early diagnosis of ABL and homozygous hypobetalipoproteinemia is extremely important as vitamin E deficiency culminates in atypical retinitis pigmentosa, spinocerebellar degeneration with ataxia, vitamin K deficiency can lead to a significant bleeding diathesis, vitamin A deficiency can contribute to eye disorders, and vitamin D deficiency can lead to defects in bone formation (1,13). High dose supplementation with fat soluble vitamins early in life can prevent or delay these devastating complications (Table 3) (7,13). Additional treatment measures include a low-fat diet and supplementation with essential fatty acids (Table 3) (7,13).

 

 Table 3. Dietary Recommendations for Abetalipoproteinemia

Fat calories

Less than 10-15% (<15 g/day) of total daily caloric requirement. Increase as tolerated.

Essential fatty acids

Ensure 2-4% daily caloric intake of EFAs (alpha-linolenic acid/linoleic acid)

Medium chain triglycerides

May prevent or treat malnutrition

Vitamin E

100-300 IU/kg/day

Vitamin A

100-400 IU/kg/day

Vitamin D

800-1200 IU/day

Vitamin K

5-35 mg/week

Derived from (1)

 

Given the very low level of atherogenic lipoproteins and lipids associated with ABL, there was interest in inhibiting MTTP therapeutically. Lomitapide is an oral MTP inhibitor that has been developed over the course of many years (12,17). In early trials, it was tested at a relatively high dose and the side effect profile was prohibitive (nausea, flatulence, and diarrhea). The more recent clinical trial program tested lower doses with drug titration in subjects with Homozygote Familial Hypercholesterolemia (HoFH) (12,17). On an intention to treat basis, LDL-C was decreased by 40% and apolipoprotein B by 39% (12). In patients who were actually taking lomitapide, LDL-C levels were reduced by 50% (12). In addition to decreasing LDL-C levels, non-HDL-C levels were decreased by 50%, Lp(a) by 15%, and triglycerides by 45% (12). Lomitapide received the same limited indication as mipomersen for adjunctive treatment of patients with HoFH (12). Besides the gastrointestinal issues already alluded to, its side effect profile includes hepatic steatosis (12). Its long-term safety has not been established.

 

PROPROTEIN CONVERTASE SUBTILISIN/KEXIN TYPE 9 (PCSK9)

 

Proprotein convertase subtilisin/ kexin type 9 (PCSK9) belongs to the proprotein convertase class of serine proteases (18-20). After synthesis, PCSK9 undergoes autocatalytic cleavage. This step is required for secretion, most likely because the prodomain functions as a chaperone and facilitates folding (18,19). PCSK9 is associated with LDL particles and the LDL-receptor (LDLR) (20). In 2003, Abifadel reported the seminal work that mapped PCSK9 as the third locus for autosomal dominant hypercholesterolemia (Familial Hypercholesterolemia- FH) (21). This finding revealed a previously unknown actor involved in cholesterol homeostasis and served to launch a series of investigations into PCSK9 biology. As it turns out, PCSK9 functions as a central regulator of plasma LDL-C concentration (18-20). It binds to the LDLR and targets it for destruction in the lysosome (18-20). Overactivity of PCSK9 results in a decrease in LDLR and an increase in LDL-C levels while decreased activity of PCSK9 results in an increase in LDLR and a decrease in LDL-C.

 

Since the discovery of gain-of-function mutations in PCSK9 as a cause of FH, investigators have also uncovered loss of function mutations of PCSK9. Loss-of-function mutations in PCSK9 are associated with low LDL-C levels and markedly reduced ASCVD (18,19). In African Americans 2.6 percent had nonsense mutations in PCSK9 that resulted in a 28 percent reduction in LDL-C and an 88 percent reduction in the risk of coronary heart disease (22). The hypolipidemia is not associated with liver abnormalities or other disorders. Interestingly, rare individuals homozygous or compound heterozygotes for loss of function mutations in PCSK9 have been reported with extremely low levels of LDL-C (~15 mg/dL), normal health and reproductive capacity, and no evidence of neurologic or cognitive dysfunction (20,23,24). Collectively, these observations served as further motivation to pursue antagonism of PCSK9 as a therapeutic target. Antagonizing PCSK9 would prolong the lifespan of LDLR, leading to significant reductions in plasma LDL-C levels. Two fully human monoclonal antibodies (alirocumab and evolocumab) targeting PCSK9 became commercially available in 2015 and inclisiran, a small interfering RNA that inhibits translation of PCSK9 is also available. Other approaches to inhibit PCSK9 are under investigation.  

 

FAMILIAL COMBINED HYPOLIPIDEMIA   

 

Familial combined hypolipidemia (FCH) is due to loss of function mutations in the gene encoding angiopoietin-like protein 3 (ANGPTL3) (25,26). ANGPTL3 inhibits various lipases, such as lipoprotein lipase and endothelial lipase (25,26). Therefore, loss of function mutations in ANGPTL3 relinquishes this inhibition increasing the activity of lipases resulting in more efficient metabolism of VLDL and HDL particles (25,26). In addition, to increasing VLDL clearance the secretion of VLDL is also decreased due to a decrease in free fatty acid flux to the liver (25). LDL clearance is increased but the mechanism remains to be fully elucidated (25). Studies have suggested that ANGPTL3 inhibition lowers LDL-C by limiting LDL particle production due to ANGPTL3 inhibition and increased endothelial lipase activity reducing VLDL-lipid content and size, generating remnant particles that are efficiently removed from the circulation rather than being further metabolized to LDL (27).

 

Clinically, FCH manifests as panhypolipidemia (decreased triglycerides, LDL-C, HDL-C, apo B, and apo A-I) (25,26,28). Interestingly, heterozygotes for certain nonsense mutations in the first exon of ANGPTL3 have moderately reduced LDL-C and triglyceride levels while compound heterozygotes have significant reductions in HDL-C as well (25,26).  Homozygosity or compound heterozygosity for other loss-of-function mutations in exon 1 of ANGPTL3 have no detectable ANGPTL3 in plasma and striking reductions of atherogenic lipoproteins with HDL particles containing only apo A-I and preß-HDL. Individuals who are heterozygous for the loss of function mutations in ANGPTL3 have significantly reduced LDL-C and triglyceride levels and a reduced risk of atherosclerosis (25,26,28).

 

A pooled analysis of cases of familial combined hypolipidemia was published 2013 (29). One hundred fifteen individuals carrying 13 different mutations in the ANGPTL3 gene (14 homozygotes, 8 compound heterozygotes, and 93 heterozygotes) and 402 controls were evaluated. Homozygotes and compound heterozygotes (two mutant alleles) had no measurable ANGPTL3 protein. In heterozygotes, ANGPTL3 was reduced by 34-88%, according to genotype. All cases (homozygotes and heterozygotes) demonstrated significantly lower concentrations of all plasma lipoproteins (except for Lp(a)) as compared to controls. Familial combined hypolipidemia is not associated with any comorbidity. In fact, the prevalence of fatty liver was the same as controls. However, ANGPTL3 deficiency is associated with a reduced risk of cardiovascular disease (25,30).

 

Recently, evinacumab, a human monoclonal antibody against ANGPTL3, was approved for the treatment of Homozygous Familial Hypercholesterolemia (12). Evinacumab decreases LDL-C levels by mechanisms independent of LDL receptor activity (12).

 

CHYLOMICRON RETENTION DISEASE

 

Chylomicron retention disease (CMRD), known also as Anderson’s disease for the individual who first described the condition in 1961, is a rare inherited lipid malabsorption syndrome (31,32). It is due to mutations in the SAR1B gene which codes for the protein SAR1b, a small GTPase, involved in intracellular protein trafficking (31). Mutations in SAR1b result in the failure of pre-chylomicrons to move from the endoplasmic reticulum to the golgi (31). This disorder usually presents in young infants with diarrhea, steatorrhea, abdominal distention, and failure to thrive, which can improve with a low-fat diet (1,31,32). Patients with CMRD demonstrate a specific autosomal recessive hypocholesterolemia that differs from other familial hypocholesterolemias. CMRD is associated with a 50% reduction in both plasma LDL-C and HDL-C with normal fasting triglyceride levels (31,32). Mutations in SAR1B do not affect VLDL secretion by the liver. The decrease in HDL-C is postulated to be due to a decrease in Apo A-I secretion and cholesterol efflux by the small intestine (31). The mechanism accounting for the decrease in LDL-C is not clear. The usual increase in triglycerides and chylomicron levels following a fat meal is blocked (31). The duodenal mucosa is white on endoscopy and intestinal biopsy reveals cytosolic lipid droplets and lipoprotein-sized particles in enterocytes (31). As one would expect the absorption of fat-soluble vitamins (A, D, K, and E) and essential fatty acids is impaired (31,32). Neurological and eye manifestations are milder and occur at an older age compared to abetalipoproteinemia (1). Red blood cell acanthosis is rare (1). Heterozygotes with mutations in SAR1B are unaffected.

 

Treatment for individuals with CMRD is similar to that described above for individuals with ABL (32).

 

ACKNOWLEDGEMENTS

 

This work was supported by grants from the Northern California Institute for Research and Education.

 

REFERENCES

 

  1. Bredefeld C, Hussain MM, Averna M, Black DD, Brin MF, Burnett JR, Charriere S, Cuerq C, Davidson NO, Deckelbaum RJ, Goldberg IJ, Granot E, Hegele RA, Ishibashi S, Karmally W, Levy E, Moulin P, Okazaki H, Poinsot P, Rader DJ, Takahashi M, Tarugi P, Traber MG, Di Filippo M, Peretti N. Guidance for the diagnosis and treatment of hypolipidemia disorders. J Clin Lipidol 2022; 16:797-812
  2. Rimbert A, Vanhoye X, Coulibaly D, Marrec M, Pichelin M, Charriere S, Peretti N, Valero R, Wargny M, Carrie A, Lindenbaum P, Deleuze JF, Genin E, Redon R, Rollat-Farnier PA, Goxe D, Degraef G, Marmontel O, Divry E, Bigot-Corbel E, Moulin P, Cariou B, Di Filippo M. Phenotypic Differences Between Polygenic and Monogenic Hypobetalipoproteinemia. Arterioscler Thromb Vasc Biol 2021; 41:e63-e71
  3. Blanco-Vaca F, Martin-Campos JM, Beteta-Vicente A, Canyelles M, Martinez S, Roig R, Farre N, Julve J, Tondo M. Molecular analysis of APOB, SAR1B, ANGPTL3, and MTTP in patients with primary hypocholesterolemia in a clinical laboratory setting: Evidence supporting polygenicity in mutation-negative patients. Atherosclerosis2019; 283:52-60
  4. Balder JW, Rimbert A, Zhang X, Viel M, Kanninga R, van Dijk F, Lansberg P, Sinke R, Kuivenhoven JA. Genetics, Lifestyle, and Low-Density Lipoprotein Cholesterol in Young and Apparently Healthy Women. Circulation 2018; 137:820-831
  5. Bredefeld C, Peretti N, Hussain MM, Medical Advisory P. New Classification and Management of Abetalipoproteinemia and Related Disorders. Gastroenterology 2021; 160:1912-1916
  6. Burnett JR, Hooper AJ, Hegele RA. APOB-Related Familial Hypobetalipoproteinemia. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Mirzaa G, Amemiya A, eds. GeneReviews((R)). Seattle (WA) 2021.
  7. Hooper AJ, Burnett JR. Update on primary hypobetalipoproteinemia. Curr Atheroscler Rep 2014; 16:423
  8. Hooper AJ, van Bockxmeer FM, Burnett JR. Monogenic hypocholesterolaemic lipid disorders and apolipoprotein B metabolism. Crit Rev Clin Lab Sci 2005; 42:515-545
  9. Welty FK. Hypobetalipoproteinemia and abetalipoproteinemia: liver disease and cardiovascular disease. Curr Opin Lipidol 2020; 31:49-55
  10. Peloso GM, Nomura A, Khera AV, Chaffin M, Won HH, Ardissino D, Danesh J, Schunkert H, Wilson JG, Samani N, Erdmann J, McPherson R, Watkins H, Saleheen D, McCarthy S, Teslovich TM, Leader JB, Lester Kirchner H, Marrugat J, Nohara A, Kawashiri MA, Tada H, Dewey FE, Carey DJ, Baras A, Kathiresan S. Rare Protein-Truncating Variants in APOB, Lower Low-Density Lipoprotein Cholesterol, and Protection Against Coronary Heart Disease. Circ Genom Precis Med 2019; 12:e002376
  11. Crooke ST, Geary RS. Clinical pharmacological properties of mipomersen (Kynamro), a second generation antisense inhibitor of apolipoprotein B. Br J Clin Pharmacol 2013; 76:269-276
  12. Feingold KR. Cholesterol Lowering Drugs. In: Feingold KR, Anawalt B, Boyce A, Chrousos G, de Herder WW, Dhatariya K, Dungan K, Grossman A, Hershman JM, Hofland J, Kalra S, Kaltsas G, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, McGee EA, McLachlan R, Morley JE, New M, Purnell J, Sahay R, Singer F, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  13. Burnett JR, Hooper AJ, Hegele RA. Abetalipoproteinemia. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Mirzaa G, Amemiya A, eds. GeneReviews((R)). Seattle (WA)2018.
  14. Lee J, Hegele RA. Abetalipoproteinemia and homozygous hypobetalipoproteinemia: a framework for diagnosis and management. J Inherit Metab Dis 2014; 37:333-339
  15. Wetterau JR, Lin MC, Jamil H. Microsomal triglyceride transfer protein. Biochim Biophys Acta 1997; 1345:136-150
  16. Hussain MM, Shi J, Dreizen P. Microsomal triglyceride transfer protein and its role in apoB-lipoprotein assembly. J Lipid Res 2003; 44:22-32
  17. Cuchel M, Rader DJ. Microsomal transfer protein inhibition in humans. Curr Opin Lipidol 2013; 24:246-250
  18. Debose-Boyd RA, Horton JD. Opening up new fronts in the fight against cholesterol. Elife 2013; 2:e00663
  19. Seidah NG, Awan Z, Chretien M, Mbikay M. PCSK9: a key modulator of cardiovascular health. Circ Res 2014; 114:1022-1036
  20. Shapiro MD, Tavori H, Fazio S. PCSK9: From Basic Science Discoveries to Clinical Trials. Circ Res 2018; 122:1420-1438
  21. Abifadel M, Varret M, Rabes JP, Allard D, Ouguerram K, Devillers M, Cruaud C, Benjannet S, Wickham L, Erlich D, Derre A, Villeger L, Farnier M, Beucler I, Bruckert E, Chambaz J, Chanu B, Lecerf JM, Luc G, Moulin P, Weissenbach J, Prat A, Krempf M, Junien C, Seidah NG, Boileau C. Mutations in PCSK9 cause autosomal dominant hypercholesterolemia. Nat Genet 2003; 34:154-156
  22. 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
  23. 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
  24. Cariou B, Ouguerram K, Zair Y, Guerois R, Langhi C, Kourimate S, Benoit I, Le May C, Gayet C, Belabbas K, Dufernez F, Chetiveaux M, Tarugi P, Krempf M, Benlian P, Costet P. PCSK9 dominant negative mutant results in increased LDL catabolic rate and familial hypobetalipoproteinemia. Arterioscler Thromb Vasc Biol 2009; 29:2191-2197
  25. Arca M, D'Erasmo L, Minicocci I. Familial combined hypolipidemia: angiopoietin-like protein-3 deficiency. Curr Opin Lipidol 2020; 31:41-48
  26. Kersten S. Angiopoietin-like 3 in lipoprotein metabolism. Nat Rev Endocrinol 2017; 13:731-739
  27. 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 Res2020; 61:1271-1286
  28. Burnett JR, Hooper AJ, Hegele RA. Familial Combined Hypolipidemia. In: Adam MP, Feldman J, Mirzaa GM, Pagon RA, Wallace SE, Bean LJH, Gripp KW, Amemiya A, eds. GeneReviews((R)). Seattle (WA) 2023.
  29. Minicocci I, Santini S, Cantisani V, Stitziel N, Kathiresan S, Arroyo JA, Marti G, Pisciotta L, Noto D, Cefalu AB, Maranghi M, Labbadia G, Pigna G, Pannozzo F, Ceci F, Ciociola E, Bertolini S, Calandra S, Tarugi P, Averna M, Arca M. Clinical characteristics and plasma lipids in subjects with familial combined hypolipidemia: a pooled analysis. J Lipid Res 2013; 54:3481-3490
  30. Dewey FE, Gusarova V, Dunbar RL, O'Dushlaine C, Schurmann C, Gottesman O, McCarthy S, Van Hout CV, Bruse S, Dansky HM, Leader JB, Murray MF, Ritchie MD, Kirchner HL, Habegger L, Lopez A, Penn J, Zhao A, Shao W, Stahl N, Murphy AJ, Hamon S, Bouzelmat A, Zhang R, Shumel B, Pordy R, Gipe D, Herman GA, Sheu WHH, Lee IT, Liang KW, Guo X, Rotter JI, Chen YI, Kraus WE, Shah SH, Damrauer S, Small A, Rader DJ, Wulff AB, Nordestgaard BG, Tybjaerg-Hansen A, van den Hoek AM, Princen HMG, Ledbetter DH, Carey DJ, Overton JD, Reid JG, Sasiela WJ, Banerjee P, Shuldiner AR, Borecki IB, Teslovich TM, Yancopoulos GD, Mellis SJ, Gromada J, Baras A. Genetic and Pharmacologic Inactivation of ANGPTL3 and Cardiovascular Disease. N Engl J Med 2017; 377:211-221
  31. Levy E, Poinsot P, Spahis S. Chylomicron retention disease: genetics, biochemistry, and clinical spectrum. Curr Opin Lipidol 2019; 30:134-139
  32. Sissaoui S, Cochet M, Poinsot P, Bordat C, Collardeau-Frachon S, Lachaux A, Lacaille F, Peretti N. Lipids Responsible for Intestinal or Hepatic Disorder: When to Suspect a Familial Intestinal Hypocholesterolemia? J Pediatr Gastroenterol Nutr 2021; 73:4-8

 

Hypertriglyceridemia: Pathophysiology, Role of Genetics, Consequences, and Treatment

ABSTRACT

 

Hypertriglyceridemia (HTG) can result from a variety of causes. Mild to moderate HTG tracks along with the metabolic syndrome, obesity and diabetes. HTG can be the result of multiple small gene variants or secondary to several diseases and drugs. Severe HTG with plasma triglyceride (TG) levels >1000-1500 mg/dL typically results from: (1) rare variants in the lipoprotein lipase (LPL) complex, where it is termed the familial chylomicronemia syndrome (FCS), and (2) the co-existence of genetic and secondary forms of HTG, termed the multifactorial chylomicronemia syndrome (MFCS), which is a much more common cause of severe HTG.  Mild to moderate HTG is associated with an increased risk of premature cardiovascular disease (CVD), while severe HTG can lead to pancreatitis as well as an increased risk of premature CVD. Appropriate management of the patient with HTG requires knowledge of the likely cause of the HTG, to prevent itscomplications.

 

PHYSIOLOGY

 

A detailed overview of lipoprotein physiology is provided in the Endotext chapter on Lipoprotein Metabolism (1).  Here we will briefly review some aspects the metabolism of the triglyceride (TG)-rich lipoproteins, very low-density lipoproteins (VLDL) and chylomicrons (CM) of particular relevance to this chapter.

 

Secretion of TG-rich Lipoproteins Into Plasma

 

TGs are transported through plasma as VLDL), which transport TGs primarily made in the liver, and as CM, which transport dietary (exogenous) fat.  VLDL secretion by the liver is regulated in several ways.  Each VLDL particle has one ApoB100 molecule, making ApoB100 availability a key determinant of the number of VLDL particles, and hence, TG secretion by the liver.  In addition to one molecule of ApoB-100, each VLDL particle contains multiple copies of other apolipoproteins, together with varied amounts of TGs, cholesteryl esters, and phospholipids.  The extent of TG synthesis is in part determined by the flux of free fatty acids (FFA) to the liver.  The addition of TG to the developing VLDL particle in the endoplasmic reticulum is mediated by the enzyme microsomal triglyceride transfer protein (MTTP).  The pool of ApoB100 in the liver is not typically regulated by its level of synthesis, which is relatively constant, but by its level of degradation, which can occur in several proteolytic pathways (2). Insulin also plays a role in the regulation of VLDL secretion -  it decreases hepatic VLDL production by limiting fatty acid influx into the liver and decreases the stability of, and promotes the posttranslational degradation of ApoB100 (3).  Recent studies have shown that ApoC-III, an Apolipoprotein thought to primarily play a role in inhibiting TG removal (see below), also is involved in the assembly and secretion of VLDL (4).  VLDL particles (containing ApoB100) also increase in plasma in the postprandial state as well as CM that contain ApoB48 (5).

 

Consumption of dietary fat results in the formation of CM by enterocytes.  Fatty acids and monoacylglycerols that result from digestion of dietary TGs by acid and pancreatic lipases are transported into enterocytes by mechanisms that are not completely understood.  In the enterocyte, monoacylglycerol and fatty acids are resynthesized into TGs by the action of the enzymes acyl-coenzyme A: monoacylglycerol acyltransferase and acyl-coenzyme A: diacylglycerol acyltransferase 1 and 2 (DGAT 1 and 2).  The resulting TGs are packaged with ApoB48 to form CM, a process also mediated by MTTP (6).   CM then pass into the thoracic duct from where they enter plasma and acquire additional apolipoproteins.  Of particular relevance to their clearance from plasma is the acquisition of ApoC-II and ApoC-III. 

 

Catabolism of the TG-rich Lipoproteins

 

TGs in both VLDL and CM are hydrolyzed by the lipoprotein lipase (LPL) complex.  LPL is synthesized by several tissues, including adipose tissue, skeletal muscle, and cardiac myocytes.  After secretion by adipocytes, the enzyme is transported by glycosylphosphatidylinositol-anchored high-density lipoprotein–binding protein 1 (GPIHBP1) to the luminal side of the capillary endothelium, where it becomes tethered to glycosaminoglycans (GAGs). This pool of LPL is referred to as “functional LPL”, since it is available to hydrolyze TGs in both VLDL and chylomicrons. LPL can be liberated from these GAG binding sites by heparin injection. Several other proteins, reviewed in (7), regulate LPL activity. These include ApoC-II, which activates LPL, and ApoC-III, which inhibits LPL in addition to its effect on VLDL secretion alluded to earlier. Both are produced by the liver and are present on TG-rich lipoproteins.  ApoC-III also inhibits the turnover of TG-rich lipoproteins through a hepatic clearance mechanism involving the LDL receptor/LDL receptor-related protein 1 (LDLR/LRP1) axis (8).  ApoE also is present on the TG-rich lipoproteins and plays an important role in the uptake and clearance of the remnants of the TG-rich lipoproteins that result from hydrolysis of TGs in these lipoproteins. Other activators of LPL include ApoA-IV (9), ApoA-V (10-12) and lipase maturation factor 1 (LMF1) (13, 14). In addition, several members of the angiopoietin-like (ANGPTL) protein family play a role in regulating LPL activity.  ANGPTL3 is produced by the liver and is an endocrine regulator by inhibiting LPL in peripheral tissues (7, 15, 16).  ANGPTL4 is produced in several tissues (7), where it inhibits LPL in a paracrine fashion (7, 17). Both ANPGTL3 and ANGPTL4 delay the clearance of the TG-rich lipoproteins (7).

 

The core TGs in VLDL and chylomicrons are hydrolyzed by ApoC-II activated LPL; FFA thus formed are taken up by adipocytes and re-incorporated into TGs for storage, or in skeletal and cardiac muscle, utilized for energy. Hydrolysis of chylomicron- and VLDL-TG results in TG-poor, cholesteryl ester and ApoE-enriched particles called chylomicron and VLDL remnants, respectively, which under physiological conditions are removed by the liver by binding to LDL receptors, LDL receptor related protein, and cell surface proteoglycans (12, 18). Hepatic TG lipase and ApoA-V also are involved in the remnant clearance process (10-12, 19, 20).

 

The clearance of TGs from plasma is saturable when plasma TGs exceed ~500-700 mg/dL (21).  When removal mechanisms are saturated, additional chylomicrons and VLDL entering plasma cannot readily be removed and hence accumulate in the plasma. As a result, plasma TGs can increase dramatically, resulting in very high levels and the accumulation of chylomicrons in plasma obtained after an overnight fast. 

 

NORMAL RANGE FOR PLASMA TRIGLYCERIDES AND DEFINITION OF HYPERTRIGLYCERIDEMIA

 

Plasma TG levels reflect the TG content of multiple lipoprotein particles, primarily chylomicrons and VLDL. Fasting TG levels less than 150 mg/dL has been generally accepted as “normal” (22, 23). A non-fasting TG of 175mg/dL represents ~75th percentile of the normal range, and levels of ~400mg/dL represent the 97th percentile (24). Plasma TGs are heavily skewed to the right in the general population with a tail towards highest levels and vary depending upon the population mix (25). The full range of TG extends from 30mg/dL to 10,000mg/dL (22). TG levels are different between sexes, being higher in males than in females, and increase with age and development of other coexisting conditions such as central adiposity, metabolic syndrome, and diabetes (24). TG levels also vary between geographic areas, among people of different ethnic backgrounds, with higher levels observed in certain populations such as Mexicans and South Asians. Lower TG levels have been observed in people of African descent and African-Americans but this may be changing due to adoption of urban lifestyles (26). Because of this skewed distribution, logarithmic transformation is required to establish statistical normal ranges of TG levels.  There is no current widely accepted definition of elevated non-fasting TG levels, but some groups have utilized 175mg/dL as a cut point (27, 28). Due to high variability of TG levels, precise definitions for non-fasting levels are difficult to establish. It is worthwhile noting that post-prandial TG levels rarely exceed 400mg/dL even after a high fat challenge.

 

Normal Range Based on Risk of Complications of Hypertriglyceridemia

 

The major complications of hypertriglyceridemia (HTG) are (1) acute pancreatitis and (2) increased risk of atherosclerotic cardiovascular disease (ASCVD). These two complications occur at different levels of TGs, the risk of pancreatitis occurring at much higher TG levels than the risk of premature ASCVD and are discussed in detail later in this chapter. 

 

Normal Range According to Guidelines

 

Despite concerns regarding establishment of an upper limit of normal for TGs, most guidelines define values for HTG, often without a strong biological rationale. Definitions for the diagnosis of HTG provided in several guidelines are shown in Table 1.   

 

Cut points for HTG were first defined by the National Cholesterol Education Program Adult Treatment Panel (NCEP-ATP). The terms mild, moderate and severe have been used based on degree of TG elevation (table 1). In general, mild to moderate HTG reflects TG levels under 500mg/dL. Severe hypertriglyceridemia (sHTG) has been arbitrarily defined by different national guidelines as either TG levels ≥500 mg/dL by the American Heart Association (AHA)/American College of Cardiology (ACC), Multispecialty Cholesterol and Canadian Cardiovascular Society Guidelines (29, 30) or TG levels ≥880 mg/dL according to the European Society of Cardiology guidelines (31). The Endocrine Society has used severe HTG for 1000 to 1999 mg/dL and very severe HTG for values >2000 mg/dL (23).  

 

Table 1. Definition of Hypertriglyceridemia According to Various Clinical Guidelines

Guideline

Classification

Triglyceride Levels

NCEP/ ATP III (32)

American Heart Association (33)

National Lipid Association (34)

Normal

Borderline-high TGs

High TGs

Very high TGs

<150 mg/dL (< 1.7 mmol/L)

150-199 mg/dL (1.7-2.3 mmol/L)

200-499 mg/dL (2.3-5.6 mmol/L)

≥500 mg/dL (≥5.6 mmol/L)

The Endocrine Society (35)

Normal

Mild HTG

Moderate HTG

Severe HTG

Very severe HTG

<150 mg/dL (< 1.7 mmol/L)

150-199 mg/dL (1.7-2.3 mmol/L)

200-999 mg/dL (2.3-11.2 mmol/L)

1000-1999 mg/dL (11.2-22.4 mmol/L)

≥2000 mg/dL (≥22.4 mmol/L)

European Society of Cardiology/European Atherosclerosis Society (36)

Normal

Mild-moderate HTG

Severe HTG

<1.7 mmol/L (<150mg/dL)

>1.7-< 10mmol/L (150-880 mg/dL)

> 10 mmol/L (> 880mg/dL)

Hegele (22)

Normal

Mild to moderate

Severe

<2.0 mmol/L (<175 mg/dL)

2.0-10 mmol/L (175- 885 mg/dL)

>10 mmol/dL (>885 mg/dL)

 

In summary, establishing a precise definition of what constitutes abnormal TG values is fraught with difficulty.  An acceptable level for the prevention of pancreatitis is likely to be quite different from that at which CVD risk might be increased. The impact of HTG on CVD risk needs to be evaluated in the context of the family history of premature CVD, associated abnormalities of lipids and lipoproteins, and other CVD risk factors, particularly those associated with the metabolic syndrome (see below).

 

CAUSES AND CLASSIFICATION OF HYPERTRIGLYCERIDEMIA

 

In general, HTG has been classified as primary, when a genetic or familial basis is suspected, or secondary, where other conditions that raise TG levels can be identified. However, this classification is likely overly simplistic. It has become clear in the past decade that the spectrum of plasma TG levels, ranging from mild elevation to very severe HTG, is modulated by a multitude of genes working in concert with non-genetic secondary and environmental contributors. Thus, in the vast majority of individuals, mutations in multiple genes with interaction from non-genetic factors result in altered TG-rich lipoprotein synthesis and catabolism and subsequent HTG.

 

Historical Perspective

 

Phenotypic heterogeneity among patients with HTG has been historically defined by qualitative and quantitative differences in plasma lipoproteins. In the pre-genomic era, the Fredrickson classification of hyperlipoproteinemia was based on electrophoretic patterns of lipoprotein fractions (37). The phenotypes are distinguished based on the specific class or classes of accumulated TG-rich lipoprotein particles, including chylomicrons, VLDL and VLDL-remnants. This classification included 6 phenotypes, five of which included HTG in their definition (except for Frederickson type 2 A hyperlipoproteinemia, which equates with genetic primary hypercholesterolemia). It has now become apparent that except for type 1 hyperlipoproteinemia (FCS), the HTG phenotypes, particularly Frederickson type 4 and type 5 hyperlipoproteinemia, are due to the accumulation of polygenic traits predisposing to HTG. However, this classification system is dated, has neither improved clinical or scientific insight, and therefore does not find wide use at this time (22).

 

In 1973, Goldstein and colleagues characterized a variable pattern of lipid abnormalities in families of survivors of myocardial infarction that they termed familial combined hyperlipidemia (FCHL) (38).  At the same time, this phenotype of mixed or combined hyperlipidemia was observed in another cohort, where it was called multiple-type familial hyperlipoproteinemia (39).  Affected family members can present with hypercholesterolemia alone, HTG alone, or with elevations in both TGs and LDL. This pattern was estimated to have a population prevalence of 1-2% (40), making it the most common inherited form of dyslipidemia.

 

In the aforementioned study, a pattern of isolated HTG, historically called familial HTG (FHTG) also was described (38). This condition was characterized by increased TG synthesis, with secretion of normal numbers of large TG-enriched VLDL particles (41), elevated VLDL levels, but normal levels of LDL and HDL cholesterol (42).  FHTG did not appear to be associated with an increased risk of premature CVD in an early study (43), but baseline TG levels predicted subsequent CVD mortality after 20 years of follow up among relatives in families classified as having FHTG (44, 45).  

 

FCHL and FHTG were initially believed to be monogenic disorders (38). However, more recent genetic characterization of individuals with familial forms of HTG indicates that these are not disorders associated with variation within a single gene, but rather polymorphisms in multiple genes associated with HTG, as detailed below. Therefore, classification of FCHL and FHTG is potentially misleading. Nevertheless, it is important to note that FCHL as originally described is associated with a very high prevalence of premature CVD (43, 44, 46).   

 

Genetic Forms of Hypertriglyceridemia

 

It is now evident that clinically relevant abnormalities of plasma TG levels appear to require a polygenic foundation of common or rare genetic variants (22).  Common small-effect gene variants confer a background predisposition that interact with rare large-effect heterozygous variants in genes that govern synthesis or catabolism of TG-rich lipoproteins, or nongenetic secondary factors, leading to the expression of a more severe TG phenotype (47). Recently, the most prevalent genetic feature underlying severe HTG was shown to be the polygenic accumulation of common (rather than rare) variants—more specifically, the accumulation of TG-raising alleles across multiple SNP loci (48).

 

Thus, mild to moderate hypertriglyceridemic states are complex, genetically heterogeneous disorders. Mild-to-moderate HTG is typically polygenic and results from the cumulative burden of common and rare variants in more than 30 genes, as quantified by genetic risk scores. All genetic forms can be exacerbated by non-genetic factors. Because they are a consequence of interaction between multiple susceptibility genes and lifestyle factors, individuals with moderate HTG should be considered as a single group without distinction, irrespective of concomitant lipoprotein disturbances (22). Because of the complexity of these disorders, routine genetic testing is not recommended.

 

Pathogenesis of Genetic Forms of Hypertriglyceridemia

           

Genetic forms of HTG without other lipoprotein disturbances (i.e., pure HTG) are characterized by increased TG synthesis, where normal numbers of large TG-enriched VLDL particles are secreted (41, 49-51). Reduced TG clearance also has been observed in some individuals (50-52).  Affected people have elevated VLDL levels, but normal levels of LDL, and are generally asymptomatic unless clinical CVD or severe HTG develops. 

 

A variety of metabolic defects that differ among families are associated with the combined hyperlipidemia phenotype. The characteristic lipoprotein abnormalities are increased ApoB levels and increased number of small dense LDL particles (42), a phenotype similar to that seen in the metabolic syndrome and type 2 diabetes (53). These primary defects occur due to 1) hepatic overproduction of VLDL particles (41) due to increased ApoB synthesis in the setting of disordered adipose metabolism (54, 55), insulin resistance (41, 56-58), and liver fat accumulation, and, 2) impaired clearance of ApoB containing particles (59, 60). Increased VLDL secretion results in an elevated plasma ApoB and HTG (56).  Long residence time of VLDL particles favors the formation of small dense LDL (59). An abundance of small dense LDL particles traditionally is associated with the presence of HTG; however, these LDL characteristics remain even after correction of the HTG by treatment with fibrates (61, 62).

 

In addition to Apo B abnormalities, other lipoprotein disturbances include abnormal expression of ApoA-II, ApoC-III, and PCSK9.  VLDL-TG levels in combined hyperlipidemia are modulated by ApoA-II and ApoC-III (63).  Plasma PCSK9 levels are higher in these patients, and levels correlate with TG and Apo B levels (64).

 

Visceral adiposity appears to be an important determinant of insulin resistance, which occurs commonly in subjects with both isolated HTG (65) and combined hyperlipidemia (65-69). Other abnormalities that have been reported in clinical FCHL include impaired lipolysis due to decreased cyclic AMP dependent signaling (54, 69), abnormal adipocyte TG turnover (70), fatty liver (71), increased arterial stiffness (72), and increased carotid intimal-medial thickness (73). 

 

In all of the phenotypes described above, severe HTG can occur when secondary causes of HTG such as untreated diabetes, marked weight gain, or use of TG-raising drugs are present concurrently, leading to the Multifactorial Chylomicronemia Syndrome (MFCS), described later (74).

 

Secondary Forms of Hypertriglyceridemia

 

These are described in greater detail in the chapters on Secondary Disorders of Lipid and Lipoprotein Metabolism (75-78).  However, in the section where we describe MFCS we will briefly touch on some aspects of secondary forms of HTG, since they assume importance in the pathogenesis of the severe HTG seen in the MFCS, where they often co-exist in individuals with genetic forms of HTG. In our experience, the commonest secondary forms of HTG that interact with genetic forms of HTG are type 2 diabetes (usually as part of the metabolic syndrome), obesity, recent weight gain, excessive alcohol consumption, the use of drugs that can raise TGs, and chronic kidney disease (CKD)(74, 79, 80). (Table 3)

           

Severe Hypertriglyceridemia and the Chylomicronemia Syndrome

 

In the late 1960s Fredrickson, Levy and Lees (37) classified HTG into types dependent on the pattern of lipoproteins on paper electrophoresis and the presence or absence of chylomicrons in fasting plasma. They recognized that acute pancreatitis and eruptive xanthomata occurred in the presence of chylomicronemia that accumulate in what they termed Type I and Type V hyperlipoproteinemia. Chylomicrons are present in the post-prandial state, and usually are present in fasting plasma when TG levels exceed 800 mg/dL, but absent in fasting plasma below that value (81). The term chylomicronemia syndrome was first used to describe a constellation of clinical findings such as abdominal pain, acute pancreatitis, eruptive xanthoma and lipemia retinalis that occurred in association with very high TG levels (82). Two groups of conditions can lead to severe HTG and clinical manifestations of the chylomicronemia syndrome; (1) familial chylomicronemia syndrome (FCS) due to variants in the LPL complex, and (2) multifactorial chylomicronemia syndrome (MFCS), in which genetic predisposition and secondary forms of HTG co-exist.

 

FAMILIAL CHYLOMICRONEMIA SYNDROME (FCS)

 

FCS is a monogenic disorder due to variants in one or more genes of the LPL complex that affect chylomicron catabolism. FCS incidence is very rare, with an estimated prevalence ranging from 1 in 20,000 to 300,000 (83).    

 

Genetics: Biallelic loss of function variants in five canonical genes lead to impaired hydrolysis of TG-rich lipoproteins, with subsequent increases in chylomicron particle numbers and markedly increased TG concentrations. The most common gene affected in FCS is LPL itself, in which patients are homozygous or compound heterozygous for two defective LPL alleles. Over 180 variants that result in LPL deficiency have been described with some clustered mutations due to founder effects (84-87). Loss of function variants account for over 90% of cases (83). Many are missense variants, some in catalytically important sites and some in regions that predispose to instability of the homodimeric structure of LPL required for enzyme activity (88). However, many common LPL gene variants have been described that have no clinical phenotype (89). Variants in the APOC2 gene, encoding ApoC-II, an activator of LPL, is another cause of FCS.  Variants have been described in several families (90, 91). In FCS thus far there is no known gene variant that affects synthesis or production of TG-rich lipoproteins.

 

FCS can also occur from biallelic loss of function variants in other components of the LPL complex, namely APOA5, LMF1, and GPIHBP1 genes (Table 2), each of which plays an important role in determining LPL function (92). The lipoprotein phenotype in these mimics that seen in classical LPL deficiency. Loss of function variants in GPIHBP1, which directs transendothelial LPL transport and helps anchor chylomicrons to the endothelial surface near LPL, thereby providing a platform for lipolysis, has been described in several families (83).  Autoantibodies to GPIHBP1 also can lead to chylomicronemia (93). A small number of individuals with homozygous variants in Apo A-V, which stabilizes  the lipoprotein–enzyme complex thereby enhancing lipolysis (10), have been described (94). Variants in LMF1, an endoplasmic reticulum chaperone protein required for post-translational activation of LPL, have also been identified in a few individuals (95).

 

Clinical presentation: FCS usually manifests in childhood or early adolescence with nausea, vomiting, failure to thrive and recurrent abdominal pain in infancy and childhood. Occasionally it can present in adulthood (87) but this is often due to delayed diagnosis with median age at diagnosis being due to unfamiliarity in most healthcare providers (96). Adults may report “brain fog” or transient confusion.

 

Classical clinical findings include eruptive xanthomas (often seen on buttocks, back, extensor surfaces of upper limbs), lipemia retinalis, and hepatosplenomegaly. Less common symptoms of FCS can include intestinal bleeding, anemia, and neurological features such as irritability and seizures. Patients present with TG levels ≥1000 mg/dL and often much higher, due to abnormal accumulation of chylomicrons, which can be detected by the appearance of lipemic/milky plasma.  Despite prolonged overnight fasting, plasma TG levels are >1000mg/dL due to the presence of chylomicrons in the circulation as a result of impaired clearance. The most serious concern, however, is the development of acute pancreatitis, which can lead to systemic inflammatory response syndrome, multi-organ failure, and death.

 

Table 2. Genetic Disorders Resulting in Familial Chylomicronemia Syndromes (FCS)

Disorder

Inheritance

Incidence

Lipid Phenotype

Underlying Defect

Clinical Features

LPL deficiency

Autosomal Recessive

1 in 1,000,000

Marked HTG/ chylomicronemia in infancy or childhood

Very low or absent LPL activity; circulating inhibitor of LPL

Hepato-splenomegaly; severe chylomicronemia

Apo C-II deficiency

Autosomal Recessive

Rare

Marked HTG/ chylomicronemia in infancy or childhood

Absent Apo C-II

Hepato-splenomegaly; severe chylomicronemia

Apo A-V mutation

Autosomal Recessive

Rare

Marked HTG/ chylomicronemia in adulthood

Defective or absent Apo A-V

Chylomicronemia

GPIHBP1 mutation

Autosomal Recessive

Rare

Marked HTG/ chylomicronemia in adulthood

Defective or absent GPIHBP1

Chylomicronemia

LFM1 mutation

Autosomal Recessive

Rare

Marked HTG/ chylomicronemia in adulthood

Defective or absent LFM1

Chylomicronemia

 Adapted from Ref (83)

 

MULTIFACTORIAL CHYLOMICRONEMIA SYNDROME (MFCS)

 

In contrast to FCS, MFCS is more common and complex. The prevalence of MFCS is much higher than FCS and estimated to be ~1:600-1000 (84).   

 

Genetics: MFCS has a genetic basis, but unlike FCS (where recessive or biallelic variants in the affected genes are causative), the genetic alteration does not always result in the phenotypic expression of the trait but only increases the possibility of the risk of developing the condition. Other factors, including non-genomic effects (epigenetics, methylation), gene-gene, or gene-environment interactions can also contribute. MFCS develops due to two main types of genetic factors that increase the odds that a patient will develop very high TG levels. First, heterozygous rare large-effect variants in one of the five canonical TG metabolism genes (LPL, Apo C-II, Apo A5, LMF-1 and GPIHBP-1) can contribute to TG elevations. These variants have variable penetrance, i.e.- clinical presentation can vary from normal to severe hypertriglyceridemia.

 

Secondly, the presence of a high burden of common small-effect TG-raising SNPs; cumulatively, these common SNP alleles increase susceptibility for developing hypertriglyceridemia. These SNPs may have an indirect impact of the metabolism of TG-rich

Lipoproteins. There is incomplete understanding of how an excess burden of SNPs contributes to TG levels, but their prevalence in patients with severe hypertriglyceridemia has been consistently demonstrated.

 

Several polygenic risk scores (PRS) for TG levels have been published (97). A recent study found that 32.0% of patients had a high polygenic score of TG-raising alleles across 16 loci compared to only 9.5% of normolipidemic controls (25). When the PRS is high, there is a significantly increased risk of developing HTG but this is not diagnostic or definitive.

 

Secondary Causes Contributing to Severe Hypertriglyceridemia in MFCS: The most common secondary cause in the past was undiagnosed or untreated diabetes (74), although earlier detection of diabetes may be making the association of marked hyperglycemia of untreated diabetes with very severe HTG less common. In addition, individuals with the metabolic syndrome and obesity have mild to moderate HTG which can become severe HTG; weight regain following successful weight loss can lead to marked HTG (23, 84). These patients almost always have relatives with genetic forms of HTG, whose TG levels are considerably lower than the index patient with severe HTG, in whom secondary forms of HTG also are present (74).  MFCS can result from the addition of specific drugs in patients with a genetic predisposition (23). These drugs include beta-adrenergic blocking agents (selective and non-selective) and/or diuretics (thiazides and loop-diuretics such as furosemide) used for hypertension, retinoid therapy for acne, oral estrogen therapy for menopause or birth control, selective estrogen receptor modulators (particularly raloxifene) for osteoporosis or breast cancer, protease inhibitors for HIV/AIDS, atypical anti-psychotic drugs, alcohol, and possibly sertraline (84).  Rarer causes of very severe HTG include autoimmune disease (sometimes with LPL- or GPIHBP1- specific antibodies), asparaginase therapy for acute lymphoblastic leukemia (98), (99) and bexarotene, a RXR agonist used in the treatment of cutaneous T cell lymphoma (100).  

 

Table 3. Secondary Causes That Can Contribute to Severe HTG

Conditions

Hypothyroidism

Suboptimally managed or new onset diabetes

Obesity

Sudden weight gain, weight regain after weight loss

Chronic kidney disease

Nephrotic syndrome

Pregnancy

Acute hepatitis

Sepsis

Inflammatory disorders

Cushing syndrome

Autoimmune chylomicronemia

            Systemic lupus erythematosis

            Anti-LPL antibodies

GPIHBP-1 antibodies

Rare Genetic Causes

Glycogen storage disorders

Lipodystrophies

            Congenital- generalized or partial

            Acquired- HIV, autoimmune

Drugs

Alcohol ingestion

Beta blockers

Thiazide diuretics

Oral estrogens

Selective estrogen reuptake modulators - tamoxifen, raloxifene, clomiphene

Androgens

Glucocorticoids

Atypical anti-psychotics

Sertraline

Bile acid resins

Sirolimus, tacrolimus

Cyclosporine

RXR agonists -bexarotene, isotretinoin, acetretin

HIV Protease inhibitors

L- asparaginase

Alpha-interferon

Propofol

Lipid emulsions

 

Following correction of treatable secondary forms of HTG in the MFCS, TG levels usually decrease to the moderately elevated levels seen in their affected relatives (101, 102).  

 

OTHER CONDITIONS RESULTING IN HYPERTRIGLYCERIDEMIA

 

Familial Dysbetalipoproteinemia (FDB or Remnant Removal Disease)

 

Familial dysbetalipoproteinemia, also referred to as remnant removal disease or type III hyperlipoproteinemia, is a rare autosomal recessive disorder that can present with elevated TG levels. This disorder is characterized by the accumulation of remnant lipoproteins.

 

PATHOGENESIS AND GENETICS

 

Remnant removal disease requires homozygosity for the ApoE2 genotype or a rare heterozygosity for a variant in the ApoE gene, which results in pathologic accumulation of remnant lipoproteins in the circulation due to impaired hepatic uptake of ApoE-containing lipoproteins (103). ApoE is a glycoprotein synthesized in the liver, brain and tissue macrophages and present on chylomicrons, VLDL and HDL. Apo E through interaction with the LDLR and heparan sulphate proteoglycans promotes the hepatic clearance of remnants of chylomicrons and VLDL (104); it also facilitates cholesterol efflux from macrophages to HDL (105). In humans, there are 3 common isoforms of ApoE , ApoE2, ApoE3, and ApoE4 (106).  Each differs in isoelectric point by one charge unit, ApoE4 being the most basic isoform and ApoE2 the most acidic.  ApoE3 (Cys112Arg158) is the commonest isoform.  ApoE2 (Arg158Cys) and ApoE4 (Cys112Arg) differ from ApoE3 by single amino acid substitutions at positions 158 and 112, respectively (107).  In the majority of cases (90%), remnant removal disease is associated with the E2/E2 genotype and results from impaired binding to the Apo E receptor. It is  an autosomal recessive disorder with the prevalence of ApoE2 homozygosity in Caucasian populations estimated to be about 1% (108).  Rarer ApoE variants such as ApoE3-Leiden (109) and ApoE2 (Lys1463Gln) that also can cause remnant accumulation are dominantly inherited (110) and account for 10% of cases (111, 112).  Rare APOE variants in the population, other than the APOE2 and APOE4 alleles, play an important role in the development of isolated hypercholesterolemia (113) and mixed hyperlipidemia, with and without familial dysbetalipoproteinemia (114). Thus it is becoming apparent that two different DBL phenotypes may exist- a genetic Apo E dysfunction and a multifactorial form (115). Modern prevalence of FDB is estimated at 1-2% (116).

 

In the absence of additional genetic, hormonal, or environmental factors, remnants do not accumulate to a degree sufficient to cause hyperlipidemia in ApoE2 homozygotes; in fact, lipid levels are commonly low. Remnant accumulation results when the E2/2 genotype is accompanied by a second genetic or acquired defect that causes overproduction of VLDL such as obesity or diabetes (117) (111, 118) , a decrease in remnant clearance, or a reduction in LDL receptor activity (e.g., hypothyroidism (119)). Thus, full phenotypic expression requires the presence of other environmental or genetic factors (120). In these circumstances, the reduced uptake of remnant lipoproteins by the liver results in reduced conversion of VLDL and intermediate density lipoproteins to LDL, with subsequent accumulation of remnant lipoproteins (121, 122), hence the term remnant removal disease.

 

DIAGNOSIS

 

Patients with remnant removal disease have roughly equivalent elevations in plasma cholesterol and TGs. The disease rarely manifests before adulthood, and in some individuals never manifests clinically. It is more common in men than in women, where expression seldom occurs before menopause, since estrogen has a protective effect in women who are ApoE2 homozygotes (108). Palmar xanthomas (Figure 1), orange lipid deposits in the palmar or plantar creases, are pathognomonic of remnant removal disease but are not always present (123). Tuberoeruptive xanthomas can be found at pressure sites on the elbows, knees and buttocks. The presence of remnant removal disease should be suspected when total cholesterol and TG levels range from 300 to 1000 mg/dL and are roughly equal in magnitude. Special diagnostic tests such as beta-quantification or lipoprotein electrophoresis are often required and are time consuming and not widely available. VLDL particles are cholesterol- enriched, which can be determined by isolation of VLDL by ultracentrifugation and by the demonstration of beta migrating VLDL on lipoprotein electrophoresis.  A VLDL-cholesterol/plasma TG ratio of <0.30 is usually observed (124).  A low ApoB/total cholesterol ratio of <0.33 also can be helpful in making the diagnosis (125). Simplified criteria for the diagnosis of DBL using a 3-step process has been proposed (126). The diagnosis of remnant removal disease should be confirmed by demonstrating the presence of the E2/E2 genotype. If the genotype result is not E2/E2, an autosomal dominant variant of APOE should be suspected. There is a high prevalence of premature coronary artery disease (127-129) and peripheral arterial disease (130-132).  Occasionally severe HTG and an increased risk of pancreatitis can develop in the presence of a concomitant secondary form of HTG or TG-raising drugs.

Figure 1. Palmar Xanthomas: Orange-yellow discoloration confined to the palmar creases.

 

Familial Partial Lipodystrophy (FPLD) Syndromes

 

A distinct entity that results in moderate and severe hypertriglyceridemia include partial lipodystrophy syndromes. Inherited lipodystrophies are a heterogeneous group of disorders considered to be rare, that manifest as complete or partial loss of white adipose tissue with accompanying severe metabolic dysregulation(133) and are reviewed elsewhere in the Endotext chapter on Lipodystrophies (134).  Loss of fat can be either localized to small discrete areas, in some cases partial with loss from extremities, or generalized with fat loss from nearly the entire body. Inherited lipodystrophies, while rare, can be autosomal dominant or recessive.  Some forms manifest at birth, while others become evident later in life.

 

Partial or generalized lipodystrophic disorders frequently are associated with significant metabolic derangements associated with severe insulin resistance, including HTG. The extent of fat loss sometimes determines the severity of metabolic complications (135).  HTG is a common accompaniment of many lipodystrophies, often in conjunction with low HDL-C levels.    The pathophysiology of hypertriglyceridemia in these subjects is possibly related to the reduced ability to deposit free fatty acids in adipose tissue due to its maldevelopment, and accelerated lipolysis with increased hepatic VLDL synthesis and delayed clearance (135).

 

Genetics: Several genes have been implicated in the manifestation of various forms including LMNA, PPARG, LIPE, CIDEC (136). In the Dunnigan variety, the most commonly identified genetic variant of FPLD, the commonest variants are in the LMNA gene and less frequently PPARG (133).  No specific genetic defect has been identified in Köbberling’s FPLD, although recent evidence suggests a heavy polygenic burden in these individuals (137, 138).

 

Diagnosis: Congenital generalized lipodystrophy (CGL) is a rare autosomal recessive disorder in which near total absence of subcutaneous adipose tissue is evident from birth.  HTG and hepatic steatosis are evident at a young age and are often difficult to control. Severe HTG, often associated with eruptive xanthoma and recurrent pancreatitis, can occur in patients with CGL. The prevalence of HTG in case series of CGL patients is over 70% (135, 139).  Plasma TGs are normal or slightly increased during early childhood, with severe HTG manifesting at puberty along with onset of diabetes mellitus. 

 

Familial partial lipodystrophies (FPLD) are complex metabolic disorders that are often not recognized clinically (140).  Partial lipodystrophies are characterized by partial loss of adipose tissue and significant metabolic derangements.  The Dunnigan variety of FPLD (FPLD type 2) is a rare autosomal dominant disorder in which fat loss mostly involves the extremities and the trunk.  Onset of fat loss in the buttocks and extremities occurs at puberty or late adolescence, with gain of fat to the face and neck. Acanthosis nigricans, calf muscle hypertrophy, and phlebomegaly (prominent veins) due to lack of subcutaneous fat, can be observed. Significant metabolic dysfunction including diabetes, which is often very insulin resistant, resistant hypertension, and HTG often severe and difficult to treat, can occur.  Myopathy, cardiomyopathy, and/or conduction system abnormalities can occur (141).  ASCVD risk also is increased (142, 143).

 

Some lipodystrophies, where fat loss appears to be proportionate to loss of total and lean body mass, do not result in dyslipidemia. Elevated TG levels have been reported in patients with atypical progeroid syndrome due to LMNA mutations (144, 145).  Of the acquired lipodystrophies, the HIV-associated form usually is characterized by more moderate HTG.  HIV-associated lipodystrophy occurs in patients receiving protease inhibitor containing highly active anti-retroviral therapy regimens (146).  Fat loss occurs in the face, buttocks, and extremities.

 

CONSEQUENCES OF HYPERTRIGLYCERIDEMIA

 

Atherosclerotic Cardiovascular Disease

 

EPIDEMIOLOGY

 

HTG has long been known to be a risk factor for ASCVD (33, 147-150), which has been confirmed in meta-analyses (45).  However, HTG also is frequently associated with low levels of HDL-cholesterol and an accumulation of remnants of the TG-rich lipoproteins, both known risk factors for ASCVD.  When adjusted for both HDL-C and non-HDL-C, which contains both remnants of the TG-rich lipoproteins and LDL, the association of TGs with ASCVD risk remained significant, although somewhat attenuated (151).  Postprandial TGs are elevated throughout the day in subjects with HTG, and postprandial TG-rich lipoproteins and their remnants also have been hypothesized to be important in the pathogenesis of atherosclerosis (150). It is therefore of interest that non-fasting TGs have been associated with ASCVD risk (150, 152, 153), despite non-fasting TGs being quite variable. However, unlike the situation with elevated LDL-C levels, the magnitude of the TG elevation does not appear to correlate with the extent of ASCVD risk. In particular, very severe HTG per se does not always appear to confer increased ASCVD risk, possibly because the chylomicrons that accumulate are too large to enter the arterial intima (154, 155). 

 

TRIGLYCERIDES IN THE PATHOGENESIS OF ASCVD

 

Although chylomicrons may be too large to enter the arterial intima, ApoE-and cholesterol-enriched remnants of the TG-rich lipoproteins can enter with ease (153) where they can bind to vascular proteoglycans, similar to LDL (156, 157).  Modification of these retained lipoproteins by either oxidative damage or enzyme digestion of some of the lipid components can liberate toxic by-products, which have been hypothesized to play a role in atherogenesis by facilitating local injury, generation of adhesion molecule, and cytokine expression and inflammation (157).  Remnants of the TG-rich lipoproteins also can be taken up by macrophages leading to the formation of foam cells, an important component of atherosclerotic plaques.  HTG also is associated with a preponderance of small, dense LDL, particles, reduced levels of HDL-C, and in the metabolic syndrome, with abnormalities of HDL composition (see earlier). Small, dense LDL can traverse the endothelial barrier more easily than large, buoyant LDL particles (158), are retained more avidly than large, buoyant LDL (159), and also are more readily oxidized (160, 161), all of which may facilitate atherogenesis. HDL particles in some hypertriglyceridemic states, e.g., in association with the metabolic syndrome, might be dysfunctional with respect to their cholesterol efflux, anti-inflammatory, and anti-oxidant properties. Moreover, a hypercoagulable state has been reported in association with both HTG and the metabolic syndrome (162). Thus, HTG might accelerate atherosclerosis by several mechanisms, all of which could increase CVD risk.

 

GENETIC EVIDENCE OF HYPERTRIGLYCERIDEMIA AND ATHEROSCLEROSIS

 

Recent human genetic studies have provided important insight into the contribution of TGs to ASCVD. Several genetic approaches, including candidate gene sequencing, GWAS of common DNA sequence variants, and genetic analysis of TG phenotypes have unraveled new proteins and gene variants involved in plasma TG regulation (163). Some genetic variants that influence TG levels appear to be associated with increased CVD risk even after adjusting for their effects on other lipid traits (164).  GWAS have identified common noncoding variants of the LPL gene locus associated with TG and CVD risk (165, 166).  A common gain-of-function mutation in the LPL gene, S447X (10% allele frequency), is associated with reduced TG levels and reduced risk of CVD (167) and an LPL variant associated with reduced TG and ApoB levels was associated with reduced CVD similar to LDL-C lowering variants, suggesting that the clinical benefit of lowering triglyceride and LDL-C levels may be proportional to the absolute change in ApoB (168).  Conversely, several loss-of-function LPL variants linked with elevated TG levels are associated with increased CVD risk (169). Variants in the TRIB1 locus have been associated with LDL, HDL-C and TG levels (166), hepatic steatosis (170) and coronary artery disease (171). Mutations that disrupt APOC3 gene function and reduce plasma ApoC-III concentration are associated with lower TG levels and decreased risk of clinical CVD (172, 173).  In contrast, carriers of rare mutations in APOA5, encoding ApoA-V, an activator of LPL, are associated with elevated TGs and with increased risk of myocardial infarction (174, 175).  Loss of function variants in ANGPTL4 that had lower TG levels also were associated with reduced CVD risk (176, 177).  Thus, exciting new human genetics findings have causally implicated TG and TG-rich lipoproteins in the development of CVD risk. In particular, the LPL pathway and its reciprocal regulators ApoC-III and ApoA-V appear to have an important influence on atherosclerotic CVD risk. However, despite this mountain of evidence demonstrating a causal relationship of TG with atherosclerosis, the possible involvement of a correlated trait, usually low HDL-C levels, or other unmeasured traits, cannot be ruled out.

 

CLINICAL TRIAL EVIDENCE OF TRIGLYCERIDE LOWERING AND ASCVD

 

In the pre-statin era, use of gemfibrozil monotherapy demonstrated cardiovascular benefit in men with coronary heart disease. However, since the advent of statins, drugs that specifically lower plasma triglyceride levels have not clearly been shown to have a benefit with cardiovascular risk reduction in clinical trials when added to background statin therapy. Reasons for this are unclear. In genetic studies to get a comparable reduction in Apo B and coronary heart disease (CHD) risk in clinical trials, a TG reduction of ~70mg/dL is required compared to a decrease in LDL-C of only 14mg/dL. Additionally, lipid alterations due to genetics are lifelong and result in much bigger reductions in CHD than a 5-year drug study. Based on genetic studies, the magnitude of TG reduction required to demonstrate cardiovascular benefit is quite large.

 

CARDIOVASCULAR DISEASE IN THE CHYLOMICRONEMIA SYNDROME

 

As described earlier, chylomicrons have been considered to be too large to penetrate the vascular endothelium and play a role in atherogenesis (152), although  remnants of the TG-rich lipoproteins may be atherogenic (152, 178-181). The incidence of CVD is low in individuals with FCS (182), although premature atherosclerosis has been documented in well characterized subjects with this disorder (183).  However, CVD risk clearly is increased in many patients with MFCS, although the exact frequency remains unclear. The frequency of CVD outcomes does not appear to relate to the magnitude of the TG elevations (184). It is not surprising that CVD is increased in MFCS considering the association between TGs and CVD that has been documented in many studies (reviewed in (150, 185, 186)).  Many subjects develop severe HTG due to the co-existence of polygenic mutations that result in mild to moderate HTG (22) with secondary causes of HTG.  Residual HTG due to these genetic disorders persists even after severely elevated TG levels have been reduced by treatment of the secondary forms of HTG and treatment of the HTG per se.  Moreover, many patients with the MFCS have other CVD risk factors such as diabetes, reduced levels of HDL-C, and hypertension, the latter resulting in use of diuretics and beta-blockers, which play a role in raising their TGs to levels at which chylomicrons accumulate due to saturation of clearance mechanisms. Therefore, strategies to prevent CVD need to be undertaken once the TGs have been lowered to a level where pancreatitis is unlikely to recur. 

 

Pancreatitis

 

Severe hypertriglyceridemia is the third most common cause of acute pancreatitis after alcohol and gallstones. The chylomicronemia syndrome describes a constellation of findings that occur with severe elevations of plasma TG levels. There is some lack of consensus as to what constitutes severe HTG, values >1000-1500 mg/dL are generally classified as severe, although some groups consider values in the 500-1000 mg/dL range also severe hypertriglyceridemia (187). 

 

Individuals with both FCS and MFCS often present with hypertriglyceridemia induced acute pancreatitis, which can be recurrent if triglyceride levels remain elevated persistently. Women with genetic HTG can develop severe HTG and pancreatitis during pregnancy particularly during the third trimester (188).

 

The pancreatitis that occurs with severe HTG can be recurrent.  In a prospective study of patients admitted with acute pancreatitis, the distribution of plasma TGs was bimodal when measured at the peak of the pain (101, 102).  TG levels <880 mg/dL were associated with gall bladder disease and chronic alcoholism, while those above 2000 mg/dL were associated with the simultaneous presence of familial and secondary forms of HTG.  It has been suggested that individuals become prone to the development of TG-induced pancreatitis at TG values between 1500-2000 mg/dL (189).  TG-induced pancreatitis has been reported with TG levels lower than 500 mg/dL(190, 191),  although in our experience this usually occurs when patients with severe HTG stopped eating some time prior to the blood draw. The frequency of severe HTG leading to acute pancreatitis varies widely from about 6-20% of subjects, possibly related to the type of patient presenting to different type of medical centers (192, 193).  Pancreatitis often is recurrent if HTG is not appreciated to be the cause and if TG levels are not adequately controlled (87). With long term multiple episodes of acute, recurrent pancreatitis, exocrine pancreatic insufficiency or insulin deficient secondary diabetes may occur. A meta-analysis of observational studies suggests that TG-induced pancreatitis has worse outcomes than pancreatitis from other causes, with an approximate doubling of renal and respiratory failure, a nearly 4-fold increase of shock and a near doubling of mortality (194). Pancreatitis due to very severe HTG also may occur during infusion of lipid emulsions for parenteral feeding (195) or with use of the anesthetic agent propofol, which is infused in a 10% fat emulsion (196). 

 

MECHANISM OF SEVERE HYPERTRIGLYCERIDEMIC PANCREATITIS

 

The mechanism by which very severe HTG leads to pancreatitis remains speculative. Suggested mechanisms include the local liberation of FFA from TGs and lysophosphatidylcholine from phosphatidycholine when pancreatic lipase encounters very high levels of TG-rich lipoproteins in the pancreatic capillaries (197). High local concentrations of FFA overwhelm the binding capacity of albumin with resultant aggregation into micellar structures with detergent properties.  Both FFA and lysophosphatidylcholine have been shown to cause chemical pancreatitis when infused into pancreatic arteries in animal models (198-200). This leads to local liberation of more lipases from the damaged pancreatic acini, resulting in a vicious cycle (198, 201).  It also has been hypothesized that increased plasma viscosity due to the presence of increased numbers of chylomicrons in the pancreatic microcirculation contributes to the development of pancreatitis (202). There also is recent evidence of gene associations in TG-induced pancreatitis; in a Chinese cohort with HTG, a CFTR variant and TNF alpha promoter polymorphism were found to be independent risk factors for developing pancreatitis (203), while another study found an increased frequency of ApoE4 (204).

 

DIAGNOSIS OF SEVERE HYPERTRIGLYCERIDEMIC INDUCED PANCREATITIS

 

The diagnosis of HTG-associated pancreatitis can be made by the presence of severely elevated TG levels in a patient with acute pancreatitis. Falsely low serum amylase levels can be encountered due to assay interference by the TG-rich lipoproteins (205). Pseudohyponatremia due to the presence of large numbers of TG-rich lipoproteins in plasma can be seen with very high TG levels. Interference with liver transaminase assays may also occur, giving spuriously high values making it difficult to exclude alcoholic liver disease (205).

 

With chronic chylomicronemia, patients may develop eruptive xanthomata (Figure 2). These xanthomas represent an inflammatory response to the deposition of chylomicron-associated lipids in tissues and are yellow-red papules that usually appear on the buttocks, back and extensor surfaces of the upper limbs. Histologically, these lesions contain lipid laden foamy macrophages (206).  

 

Figure 2. Eruptive Xanthomas. The commonest site is on the buttocks. The lesions are papular with an erythematous base. They often are itchy.

 

Lipemia retinalis, where the retinal vessels take on a whitish hue with pallor of the optic fundus and retina can be observed with very high TG levels (Figure 3).  There is no associated visual impairment.  

Figure 3. Lipemia retinalis. Note the pale color of the retinal vessels.

 

Acute recent memory loss and mental fogginess (82) can also occasionally be seen, but has not been extensively studied. Symptoms such as fatigue, blurred vision, dysesthesias, and transient ischemic attacks have been suggested to be related to hyperviscosity resulting from high TG levels (207, 208).  Hepatosplenomegaly is frequently present in FCS due to macrophage infiltration in response to the chylomicron accumulation. Fatty liver is a common finding on imaging in both FCS and MFCS.

 

MANAGEMENT OF SEVERE HYPERTRIGLYCERIDEMIA

 

Management of HTG by lifestyle and pharmacological means is discussed in detail in the Endotext chapters on The Effect of Diet on Cardiovascular Disease and Lipid and Lipoprotein Levels and Triglyceride Lowering Drugs (209, 210).  However, in this section we will make a few points specifically relevant to this chapter. 

 

Before initiating lifelong therapy for hypertriglyceridemia, evaluation for and treatment of reversible secondary disorders that can elevate plasma triglyceride levels is crucial. This includes appropriate management of diabetes and hypothyroidism and substituting drugs that can elevate triglyceride levels with lipid-neutral agents. Management of hypertension should include calcium channel blockers, angiotensin-converting enzyme inhibitors, angiotensin receptor blockers, and alpha-adrenergic blockers rather than beta-adrenergic blocking agents and diuretics.

 

Cardiovascular Disease Prevention

 

ASCVD risk in HTG is modulated by the presence of several other factors, including other lipoprotein abnormalities, other CVD risk factors, and family history of CVD, with some families with HTG appearing to have a greater risk of CVD than others (44). The role of TG lowering by pharmacological means remains controversial, but there is consensus that the presence of HTG imparts residual risk after LDL has been adequately lowered with statins.

 

Statins: The best clinical trial data currently available for the prevention of ASCVD in patients with HTG demonstrate that statins are likely to confer the most benefit, even though their primary mode of action is not to reduce plasma TGs, nor are they very effective in so doing (211).  In patients with elevated TG levels statins will result in a significant decrease in TG levels. Based on the results of the IMPROVE-IT trial (212), the addition of ezetimibe may be of additional benefit.

 

Fibrates: Fibrates such as gemfibrozil and fenofibrate, are PPAR-α agonists, and very effective in lowering plasma TG levels (by up to 50%).  Several studies have failed to demonstrate a benefit of fibrates on ASCVD events, either alone or in combination with statins.  However, participants in these studies were not confined to individuals with HTG.  Nonetheless, post-hoc analysis showed that subgroups of subjects who had mild HTG >200mg/dL and LDL-C <34mg/dL had a significant reduction of ASCVD events (213-216).  In addition, the Action to Control Cardiovascular Risk in Diabetes (ACCORD)-LIPID trial, which was confined to subjects with diabetes, showed a similar outcome in the subgroups with HTG, although the trial was negative for all subjects (214).  Recently a novel selective peroxisome proliferator-activated receptor α modulator, pemafibrate, that possesses unique PPARα activity and selectivity (217), was evaluated in individuals with HTG and diabetes in the Pemafibrate to Reduce Cardiovascular OutcoMes by Reducing Triglycerides IN patiENts With diabetes (PROMINENT) trial. Patients with type 2 diabetes, triglyceride level 200 to 499 mg/dL and HDL-C of </=40 mg/dL were assigned to pemafibrate or placebo. Unfortunately, pemafibrate failed to demonstrate cardiovascular benefit in patients with type 2 diabetes and mild to moderate HTG despite significantly lowering triglyceride levels (218).  Thus, addition of a fibrate to a statin for cardiovascular risk reduction cannot be recommended at this time.

 

Omega-3 fish oil: Omega-3 (n-3) fatty acids are polyunsaturated fatty acids that lower TGs. The two main n-3 fatty acids are eicosapentanoic acid (EPA) and docosahexanoic acid (DHA) which can lower VLDL secretion and are agonists of PPARa. However, their role in ASCVD prevention also has been controversial as several RCTs of various dosages of n-3 mixtures failed to demonstrate CV benefit in mild to moderate HTG (219). The Reduction of Cardiovascular Events with Icosapent Ethyl–Intervention Trial (REDUCE-IT), evaluated the addition of high dose icosapent ethyl (highly purified eicosapentanoic acid or EPA) compared to placebo in high-risk patients with mild to moderate HTG on statin therapy, demonstrated a surprising 25% lower CV risk in subjects. Notably this effect was independent of baseline TG levels and TG reduction.  (220).  Subsequently the STRENGTH (Statin Residual Risk Reduction With Epanova in High Cardiovascular Risk Patients with Hypertriglyceridemia) trial of n–3 fatty acid mixtures (EPA + DHA) failed to demonstrate cardiovascular benefit and was terminated early for futility (221). Similarly, use of an EPA/DHA mixture (1.8 g/d) in patients with a recent myocardial infarction in the OMega-3 fatty acids in Elderly with Myocardial Infarction (OMEMI) trial (222) also failed to meet its primary endpoint. Tissue EPA levels may be a contributor to the positive results in the REDUCE-IT study, as there is evidence that EPA inhibits inflammation, causes membrane stabilization, and decreases plaque volume. The REDUCE-IT trial has generated controversy due to use of mineral oil in the control group which resulted in an increase in both LDL-C and C-reactive protein (223, 224). Nonetheless, several current guidelines suggest addition of icosapent ethyl in addition to a statin for residual hypertriglyceridemia in high-risk individuals (those with known ASCVD or diabetes with additional risk factors).

.

Niacin: Niacin effectively lowers triglycerides and LDL-C while raising HDL-C. Niacin inhibits lipolysis in adipocytes, therefore decreasing the available fatty acids for VLDL synthesis. However, it has fallen out of favor for ASCVD risk reduction. Two RCTs, Atherothrombosis Intervention in Metabolic Syndrome with Low HDL/High Triglycerides: Impact on Global Health Outcomes (AIM-HIGH) and Treatment of HDL to Reduce the Incidence of Vascular Events (HPS2-THRIVE) demonstrated no benefit to the addition of niacin to statins for decreasing cardiovascular risk. However, it is important to note that neither of these trials were confined to subjects with high TG levels. Therefore, due to the lack of efficacy and potential for increasing insulin resistance, niacin is not recommended for treatment of HTG.

 

Newer Therapies for HTG

 

Apo C-III Inhibitors: ApoC-III is an endogenous inhibitor of LPL, by displacing ApoC-II an activator of LPL. This leads to inhibition of lipolysis and elevations in TG levels. Apo C-III can also increase TGs by LPL independent mechanisms. Currently, ApoC-III inhibitors include volanesorsen (a second-generation antisense oligonucleotide (ASO)), olezarsen (a third generation ASO), and ARO-APOC3 (a small interfering RNA), all of which are directed at APOC3 gene (225). Volanesorsen has been studied in individuals with FCS and MFCS and has been found to decrease TG levels by up to 70% from baseline and potentially prevent pancreatitis. However, thrombocytopenia and injection site reactions are common. This drug is therefore not approved for use by the FDA due to concerns of bleeding but is approved for use by the European Medicines Agency.  

Olezarsen is a third generation ASO for which early studies have been completed.

 

ANGPTL-3 inhibitors: ANGPTL3 is a key regulator of lipoprotein metabolism and is able to repress LPL and endothelial lipase activity, with resultant increase in TGs and TG rich lipoprotein levels (226). Homozygous loss of function variants in ANGPTL3 result in combined hypolipidemia with low TG and LDL-C levels. ANGPTL3 inhibition results in decreased LDL-C, and TGs. ANGPTL3 inhibitors include evinacumab, a monoclonal antibody, vupanorsen, an antisense oligonucleotide (ASO), and ARO-ANG3, a small-interfering ribonucleic acid (siRNA). Evinacumab has been studied in patients with sHTG. A phase 2 study of evinacumab in patients with sHTG, demonstrated significant reductions in TGs only in patients with MFCS with and without LPL mutations, but not in patients with FCS, suggesting that ANGPTL3 inhibition is dependent on presence of some LPL activity (227). In the cohort with polygenic sHTG and MCS, evinacumab 15 mg/kg IV every 4 weeks resulted in an 81.7% reduction in TGs. It should be noted that evinacumab is currently only approved for homozygous familial hypercholesterolemia.

 

Management of Severe HTG-Induced Pancreatitis

 

Because of the low frequency of severe HTG in the general population, and because only some patients with severe HTG develop pancreatitis, large random controlled clinical trials are difficult to perform and unlikely to be undertaken in the foreseeable future. Therefore, therapeutic decisions need to be based on less stringent criteria than might otherwise be desirable.  However, keeping TG levels <500 mg/dL should prevent the onset of TG-induced pancreatitis (187, 228, 229). 

 

ACUTE MANAGEMENT

 

The clinical presentation of HTG-induced pancreatitis is similar to that from other causes of acute pancreatitis and can be preceded by episodic nausea, epigastric pain radiating through to the back, and increasing heart-burn. Individuals with recurrent acute pancreatitis may present without severe elevations in pancreatic enzymes (230). The immediate goal is to lower TGs in hospitalized patients.  Management is similar to the management of non-TG induced pancreatitis, which includes cessation of all oral intake for pancreatic rest, fluid resuscitation, pain management, and management of metabolic abnormalities. TGs fall rapidly with discontinuation of oral intake, often to under 1000mg/dL with cessation of oral intake. With clinical improvement, oral diet advancement should be done slowly and cautiously in the hospital as this can result in rebound TG elevations. Supportive care as needed should be instituted for organ failure. Lipid emulsions for parenteral feeding should be avoided since their use will further delay clearance and exacerbate the HTG. If long term nutrition is required for very ill individuals who cannot eat, total parental nutrition without lipids should be utilized.   

 

Heparin: Heparin will liberate LPL into plasma from its endothelial binding sites and hence rapidly lowers TGs (231). However, it also can cause rebound HTG due to rapid degradation of released LPL (232) and increases the risk of hemorrhagic pancreatitis. Therefore, the use of heparin is not recommended (233). 

 

Insulin: The rationale for the use of an IV insulin infusion of regular insulin (in conjunction with IV glucose administration as needed) is that it can activate LPL and enhance clearance of TG-rich lipoproteins (234). Intravenous insulin can be beneficial in individuals with diabetes needing glycemic control. Its use in TG--induced pancreatitis without diabetes has been reported in several case reports (235-239), and has become widespread but it is unclear whether similar changes would have occurred simply by restricting oral intake without the use of insulin.  Regular insulin at 0.1-0.2 units/kg/hour with a separate iv dextrose infusion to prevent hypoglycemia in individuals without diabetes is often used. IV insulin can be stopped when TG drops to below 1000mg/dL. However, TGs will increase when the individual consumes an oral diet; therefore, caution should be exercised with slow advancement of the diet. In a study of chylomicronemia with uncontrolled diabetes, insulin infusion lowered TGs more rapidly than plasmapheresis (240).

 

Plasmapheresis: The use of plasmapheresis to acutely lower TGs is controversial. Plasmapheresis is extracorporeal therapy where plasma is removed and replaced; plasma is separated from the blood and discarded removing chylomicrons. Substitute fluid is replaced to maintain blood volume. The procedure is highly effective in rapidly decreasing TG levels by 85% after 1 session. However, the effect is not persistent, without evidence of long-term efficacy or mortality and morbidity benefit demonstrated. Although recommended by some (241, 242), the current evidence for the benefit of use of plasmapheresis is limited to small uncontrolled anecdotal series (243) from which no firm conclusion can be made regarding its use in acute TG-induced pancreatitis (244). A recent retrospective analysis demonstrated no benefit on length of hospital stay or mortality when therapeutic plasma exchange was added to medical management of severely elevated plasma TGs (245). TG levels fall rapidly with cessation of oral intake and use of non-lipid-containing intravenous fluids.  Plasmapheresis requires a specialized center, needs central venous access, and transient anticoagulation; it only temporarily improves TG levels without addressing the underlying cause (83). Risks include line sepsis, deep vein thrombosis, and bleeding. Therefore, we do not recommend its routine use in this situation unless clinical circumstances necessitate plasmapheresis such as severe acute necrotizing pancreatitis (246), shock, or pregnancy (247). 

 

LONG-TERM MANAGEMENT TO PREVENT PANCREATITIS

 

After TG lowering in the setting of acute pancreatitis, it is essential to determine both the primary and secondary causes of the severe HTG that precipitated the acute pancreatitis.   Continued management of any secondary form of HTG, as well as lifestyle and drug therapy to maintain low TG levels is required to prevent recurrent pancreatitis. If fasting plasma TG levels remain above 1000 mg/dL after treating or removing the precipitating causes of the severe HTG, life-long therapy with fibrates or n-3 fatty acids, as described earlier, might be considered for these patients. Limited evidence suggests that orlistat, a gastrointestinal lipase inhibitor that decreases absorption of ingested fat, thereby reducing intestinal chylomicron synthesis, may be of benefit in reducing TG levels when used in conjunction with fibrate therapy (248, 249). TG and glucose control can be particularly challenging in individuals with familial partial lipodystrophy.

 

Management of Specific Syndromes that Accompany Severe HTG

 

FAMILIAL CHYLOMICRONEMIA SYNDROME (FCS)

 

Treatment of FCS includes management of an acute crisis (pancreatitis) and long-term management of HTG. Management of acute HTG-induced pancreatitis is described in the previous section. Long-term management of individuals with FCS involves patient education and maintaining a very low-fat diet. Consumption of even small amounts of fat can lead to severe HTG in FCS due to the absence of functional LPL. Infants with FCS presenting with abdominal pain or failure to thrive require discontinuation of breast feeding with replacement by very low-fat formula feeding to decrease TG levels and symptoms. In children and adults with FCS, dietary fat calories should be severely restricted to control the severe HTG and abdominal pain. This translates to about 5% to 10% of total daily calories from fat, which is a major burden for these patients (250).  Medium-chain TGs, which are taken up directly by the liver after absorption and do not enter plasma as chylomicrons via the thoracic duct, are a potential alternate fat source for these patients. n-3 fatty acids can aggravate the severe HTG of FCS and therefore are contraindicated in these individuals (251, 252).  Fibrates are not efficacious in FCS (253).  There are limited studies showing that orlistat might be beneficial in patients with FCS (254, 255).  Alcohol, oral contraceptives, and other TG-elevating drugs (see Table 3) can exacerbate severe HTG and precipitate acute pancreatitis in FCS.  Successful pregnancies in patients with FCS have become more common of late (256, 257).

 

Alipogene tiparvovec, an adeno-associated virus LPL gene therapy that was developed and resulted in significant improvement in postprandial chylomicron metabolism in patients with FCS has been abandoned and no longer available (258), Antisense oligonucleotide inhibitors of ApoC-III (volanesorsen is approved in Europe but no in the US) and of ANGPLT3 are in development for FCS (96).

 

MULTIFACTORIAL CHYLOMICRONEMIA SYNDROME (MFCS)

 

Management of acute HTG-induced pancreatitis is described in the previous section.

 

Long term management: To prevent acute HTG-induced pancreatitis in MFCS, the goal is to maintain TG levels below the threshold for pancreatitis, preferably <500 mg/dL. This requires instituting lifestyle adjustments, reversal of any secondary causes of HTG, such as treatment of suboptimally managed or undiagnosed diabetes, treating hypertension with lipid neutral agents such as ACE inhibitors, ARBs, calcium channel inhibitors, or alpha blockers rather than beta-adrenergic blockers and diuretics, and discontinuing other TG-raising drugs (table 3) where possible.  Alcohol intake should be limited or eliminated, since even small amounts of alcohol can substantially raise TG levels in individuals with baseline HTG.  Attention should be paid to avoid rebound weight gain that commonly occurs after successful weight loss. Oral estrogens should be substituted by transdermal or vaginal preparations, which raise plasma TGs to a lesser extent than oral estrogens (259, 260). Residual HTG should be treated with fibrates (229), which together with management of the secondary disorder or disorders, can reduce TG levels to below the threshold for developing pancreatitis. Other agents that can be used to lower TGs alone or in combination with fibrates, include n-3 fatty acids, and high-dose statins. We do not recommend using niacin due to risk of worsening insulin resistance and lack of clinical trial data for benefit. Lifestyle measures and weight loss are important, but patients should be educated on risks of rapid weight regain after successful weight loss can be associated with rebound severe HTG. Bariatric surgery also has been used to reduce severe HTG in refractory HTG (261). Inhibition of ApoC-III or ANGPTL3, which have been shown to lower TGs in patients with severe HTG (262), may have a role to play in the future the management of severe HTG in patients with MFCS. 

 

REFERENCES

 

  1. Feingold, K.R., Introduction to Lipids and Lipoproteins, in Endotext, K.R. Feingold, et al., Editors. 2021: South Dartmouth (MA).
  2. Fisher, E.A., The degradation of Apolipoprotein B100: multiple opportunities to regulate VLDL triglyceride production by different proteolytic pathways. Biochim Biophys Acta, 2012. 1821(5): p. 778-81.
  3. Sundaram, M. and Z. Yao, Recent progress in understanding protein and lipid factors affecting hepatic VLDL assembly and secretion. Nutr Metab (Lond), 2010. 7: p. 35.
  4. Yao, Z., Human Apolipoprotein C-III - a new intrahepatic protein factor promoting assembly and secretion of very low density lipoproteins. Cardiovasc Hematol Disord Drug Targets, 2012. 12(2): p. 133-40.
  5. Schneeman, B.O., et al., Relationships between the responses of triglyceride-rich lipoproteins in blood plasma containing Apolipoproteins B-48 and B-100 to a fat-containing meal in normolipidemic humans. Proc Natl Acad Sci U S A, 1993. 90(5): p. 2069-73.
  6. Kindel, T., D.M. Lee, and P. Tso, The mechanism of the formation and secretion of chylomicrons. Atheroscler Suppl, 2010. 11(1): p. 11-6.
  7. Kersten, S., Physiological regulation of lipoprotein lipase. Biochim Biophys Acta, 2014. 1841(7): p. 919-33.
  8. Gordts, P.L., et al., ApoC-III inhibits clearance of triglyceride-rich lipoproteins through LDL family receptors. J Clin Invest, 2016. 126(8): p. 2855-66.
  9. Goldberg, I.J., et al., Lipoprotein ApoC-II activation of lipoprotein lipase. Modulation by Apolipoprotein A-IV. J Biol Chem, 1990. 265(8): p. 4266-72.
  10. Nilsson, S.K., et al., Apolipoprotein A-V; a potent triglyceride reducer. Atherosclerosis, 2011. 219(1): p. 15-21.
  11. Priore Oliva, C., et al., Inherited Apolipoprotein A-V deficiency in severe hypertriglyceridemia. Arterioscler Thromb Vasc Biol, 2005. 25(2): p. 411-7.
  12. Gonzales, J.C., et al., Apolipoproteins E and AV mediate lipoprotein clearance by hepatic proteoglycans. J Clin Invest, 2013. 123(6): p. 2742-51.
  13. Kroupa, O., et al., Linking nutritional regulation of Angptl4, Gpihbp1, and Lmf1 to lipoprotein lipase activity in rodent adipose tissue. BMC Physiol, 2012. 12: p. 13.
  14. Lamiquiz-Moneo, I., et al., (Identification of variants in LMF1 gene associated with primary hypertriglyceridemia). Clin Investig Arterioscler, 2015.
  15. Inukai, K., et al., ANGPTL3 is increased in both insulin-deficient and -resistant diabetic states. Biochem Biophys Res Commun, 2004. 317(4): p. 1075-9.
  16. Shimamura, M., et al., Leptin and insulin down-regulate angiopoietin-like protein 3, a plasma triglyceride-increasing factor. Biochem Biophys Res Commun, 2004. 322(3): p. 1080-5.
  17. Koster, A., et al., Transgenic angiopoietin-like (angptl)4 overexpression and targeted disruption of angptl4 and angptl3: regulation of triglyceride metabolism. Endocrinology, 2005. 146(11): p. 4943-50.
  18. Foley, E.M., et al., Hepatic remnant lipoprotein clearance by heparan sulfate proteoglycans and low-density lipoprotein receptors depend on dietary conditions in mice. Arterioscler Thromb Vasc Biol, 2013. 33(9): p. 2065-74.
  19. Crawford, S.E. and J. Borensztajn, Plasma clearance and liver uptake of chylomicron remnants generated by hepatic lipase lipolysis: evidence for a lactoferrin-sensitive and Apolipoprotein E-independent pathway. J Lipid Res, 1999. 40(5): p. 797-805.
  20. Dichek, H.L., et al., Hepatic lipase overexpression lowers remnant and LDL levels by a noncatalytic mechanism in LDL receptor-deficient mice. J Lipid Res, 2001. 42(2): p. 201-10.
  21. Brunzell, J.D., et al., Evidence for a common saturable triglyceride removal mechanism for chylomicrons and very low density lipoproteins in man. Journal of Clinical Investigation, 1973. 52: p. 1578-1585.
  22. Hegele, R.A., et al., The polygenic nature of hypertriglyceridaemia: implications for definition, diagnosis, and management. Lancet Diabetes Endocrinol, 2014. 2(8): p. 655-66.
  23. Berglund, L., et al., Evaluation and treatment of hypertriglyceridemia: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab, 2012. 97(9): p. 2969-89.
  24. Subramanian, S., Approach to the Patient With Moderate Hypertriglyceridemia. J Clin Endocrinol Metab, 2022. 107(6): p. 1686-1697.
  25. Dron, J.S. and R.A. Hegele, Genetics of Hypertriglyceridemia. Front Endocrinol (Lausanne), 2020. 11: p. 455.
  26. Noubiap, J.J., et al., Prevalence of dyslipidaemia among adults in Africa: a systematic review and meta-analysis. Lancet Glob Health, 2018. 6(9): p. e998-e1007.
  27. Laufs, U., et al., Clinical review on triglycerides. Eur Heart J, 2020. 41(1): p. 99-109c.
  28. Pedersen, S.B., A. Langsted, and B.G. Nordestgaard, Nonfasting Mild-to-Moderate Hypertriglyceridemia and Risk of Acute Pancreatitis. JAMA Intern Med, 2016. 176(12): p. 1834-1842.
  29. Grundy, S.M., et al., 2018 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. Circulation, 2019. 139(25): p. e1082-e1143.
  30. Pearson, G.J., et al., 2021 Canadian Cardiovascular Society Guidelines for the Management of Dyslipidemia for the Prevention of Cardiovascular Disease in Adults. Can J Cardiol, 2021. 37(8): p. 1129-1150.
  31. Mach, F., et al., 2019 ESC/EAS Guidelines for the management of dyslipidaemias: lipid modification to reduce cardiovascular risk. Eur Heart J, 2020. 41(1): p. 111-188.
  32. Expert Panel on Detection, E. and A. Treatment of High Blood Cholesterol in, Executive Summary of The Third Report of The National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, And Treatment of High Blood Cholesterol In Adults (Adult Treatment Panel III). JAMA, 2001. 285(19): p. 2486-97.
  33. Miller, M., et al., Triglycerides and cardiovascular disease: a scientific statement from the American Heart Association. Circulation, 2011. 123(20): p. 2292-333.
  34. Jacobson, T.A., et al., National lipid association recommendations for patient-centered management of dyslipidemia: part 1--full report. J Clin Lipidol, 2015. 9(2): p. 129-69.
  35. Berglund, L., et al., Evaluation and treatment of hypertriglyceridemia: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab, 2012. 97(9): p. 2969-89.
  36. Authors/Task Force, M., et al., 2016 ESC/EAS Guidelines for the Management of Dyslipidaemias: The Task Force for the Management of Dyslipidaemias of the European Society of Cardiology (ESC) and European Atherosclerosis Society (EAS) Developed with the special contribution of the European Assocciation for Cardiovascular Prevention & Rehabilitation (EACPR). Atherosclerosis, 2016. 253: p. 281-344.
  37. Fredrickson, D., R. Levy, and R. Lees, Fat transport and lipoproteins - an integrated approach to mechanisms and disorders. N Engl J Med, 1967. 276: p. 32,94,148,215,273.
  38. Goldstein, J.L., et al., 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(7): p. 1544-68.
  39. Nikkila, E.A. and A. Aro, Family study of serum lipids and lipoproteins in coronary heart-disease. Lancet, 1973. 1(7810): p. 954-9.
  40. Brunzell, J.D., Clinical practice. Hypertriglyceridemia. N Engl J Med, 2007. 357(10): p. 1009-17.
  41. Chait, A., J.J. Albers, and J.D. Brunzell, Very low density lipoprotein overproduction in genetic forms of hypertriglyceridemia. European Journal of Clinical Investigation, 1980. 10: p. 17-22.
  42. Brunzell, J.D., et al., Plasma lipoproteins in familial combined hyperlipidemia and monogenic familial hypertriglyceridemia. Journal of Lipid Research, 1983. 24: p. 147-155.
  43. Brunzell, J.D., et al., Myocardial infarction in the familial forms of hypertriglyceridemia. Metabolism: Clinical and Experimental, 1976. 25: p. 313-320.
  44. Austin, M.A., et al., Cardiovascular disease mortality in familial forms of hypertriglyceridemia: A 20-year prospective study. Circulation, 2000. 101: p. 2777-2782.
  45. Hokanson, J.E. and M.A. Austin, Plasma triglyceride level is a risk factor for cardiovascular disease independent of high-density lipoprotein cholesterol level: a meta- analysis of population-based prospective studies. J Cardiovasc Risk, 1996. 3(2): p. 213-9.
  46. McNeely, M., et al., Lipoprotein and Apolipoprotein abnormalities in familial combined hyperlipidemia: a 20-year prospective study. Atherosclerosis, 2001. 159: p. 417-481.
  47. Lewis, G.F., C. Xiao, and R.A. Hegele, Hypertriglyceridemia in the genomic era: a new paradigm. Endocr Rev, 2015. 36(1): p. 131-47.
  48. Dron, J.S., et al., Severe hypertriglyceridemia is primarily polygenic. J Clin Lipidol, 2019. 13(1): p. 80-88.
  49. Eaton, R.P., R.C. Allen, and D.S. Schade, Overproduction of a kinetic subclass of VLDL-ApoB, and direct catabolism of VLDL-ApoB in human endogenous hypertriglyceridemia: an analytical model solution of tracer data. J Lipid Res, 1983. 24(10): p. 1291-303.
  50. Grundy, S.M., et al., Transport of very low density lipoprotein triglycerides in varying degrees of obesity and hypertriglyceridemia. Journal of Clinical Investigation, 1979. 63: p. 1274-1283.
  51. Beil, U., et al., Triglyceride and cholesterol metabolism in primary hypertriglyceridemia. Arteriosclerosis, 1982. 2(1): p. 44-57.
  52. Sigurdsson, G., A. Nicoll, and B. Lewis, Metabolism of very low density lipoproteins in hyperlipidaemia: studies of Apolipoprotein B kinetics in man. Eur J Clin Invest, 1976. 6(2): p. 167-77.
  53. Ayyobi, A.F. and J.D. Brunzell, Lipoprotein distribution in the metabolic syndrome, type 2 diabetes mellitus, and familial combined hyperlipidemia. Am J Cardiol, 2003. 92(4A): p. 27J-33J.
  54. Reynisdottir, S., et al., Impaired activation of adipocyte lipolysis in familial combined hyperlipidemia. J Clin Invest, 1995. 95(5): p. 2161-9.
  55. Reynisdottir, S., et al., Adipose tissue lipoprotein lipase and hormone-sensitive lipase. Contrasting findings in familial combined hyperlipidemia and insulin resistance syndrome. Arterioscler Thromb Vasc Biol, 1997. 17(10): p. 2287-92.
  56. Venkatesan, S., et al., Stable isotopes show a direct relation between VLDL ApoB overproduction and serum triglyceride levels and indicate a metabolically and biochemically coherent basis for familial combined hyperlipidemia. Arteriosclerosis and Thrombosis, 1993. 13: p. 1110-1118.
  57. Aitman, T., et al., Defects of insulin action on fatty acid and carbohydrate metabolism in familial combined hyperlipidemia. Arterioscler Thromb Vasc Biol, 1997. 17: p. 748-754.
  58. Janus, E.D., et al., Kinetic bases of the primary hyperlipidemias:Studies of Apolipoprotein B turnover in genetically defined subjects. European Journal of Clinical Investigation, 1980. 10: p. 161-172.
  59. Berneis, K.K. and R.M. Krauss, Metabolic origins and clinical significance of LDL heterogeneity. J Lipid Res, 2002. 43(9): p. 1363-79.
  60. Cabezas, M.C., et al., Impaired chylomicron remnant clearance in familial combined hyperlipidemia. Arterioscler Thromb, 1993. 13(6): p. 804-14.
  61. Hokanson, J.E., et al., Plasma triglyceride and LDL heterogeneity in familial combined hyperlipidemia. Arteriosclerosis and Thrombosis, 1993. 13: p. 427-434.
  62. Hokanson, J.E., et al., LDL physical and chemical properties in familial combined hyperlipidemia. Arteriosclerosis, Thrombosis, and Vascular Biology, 1995. 15: p. 452-459.
  63. Cruz-Bautista, I., et al., Determinants of VLDL composition and Apo B-containing particles in familial combined hyperlipidemia. Clin Chim Acta, 2015. 438: p. 160-5.
  64. Brouwers, M.C. and M.M. van Greevenbroek, Lipid metabolism: the significance of plasma proprotein convertase subtilisin kexin type 9 in the elucidation of complex lipid disorders. Curr Opin Lipidol, 2011. 22(4): p. 317-8.
  65. Hopkins, P.N., et al., Coronary artery disease risk in familial combined hyperlipidemia and familial hypertriglyceridemia: a case-control comparison from the National Heart, Lung, and Blood Institute Family Heart Study. Circulation, 2003. 108(5): p. 519-23.
  66. Purnell, J.Q., et al., Relationship of insulin sensitivity and ApoB levels to intra-abdominal fat in subjects with familial combined hyperlipidemia. Arterioscler Thromb Vasc Biol, 2001. 21(4): p. 567-72.
  67. Ascaso, J., et al., Insulin resistance in patients wtih familial combined hyperlipidemia and coronary artery disease. Am J Cardiol, 1997. 80: p. 1481-1487.
  68. Castro Cabezas, M., et al., Impaired fatty acid metabolism in familial combined hyperlipidemia: a mechanism associating hepatic Apolipoprotein B overproduction and insulin resistance. J Clin Invest, 1993. 92: p. 160-168.
  69. van der Kallen, C., et al., Evidence of insulin resistant lipid metabolism in adipose tissue in familial combined hyperlipidemia, but not type 2 diabetes mellitus. Atherosclerosis, 2002. 164: p. 337-346.
  70. Arner, P., et al., Dynamics of human adipose lipid turnover in health and metabolic disease. Nature, 2011. 478(7367): p. 110-3.
  71. Brouwers, M.C., et al., Fatty liver is an integral feature of familial combined hyperlipidaemia: relationship with fat distribution and plasma lipids. Clin Sci (Lond), 2007. 112(2): p. 123-30.
  72. Brouwers, M.C., et al., Increased arterial stiffness in familial combined hyperlipidemia. J Hypertens, 2009. 27(5): p. 1009-16.
  73. Keulen, E.T., et al., Increased intima-media thickness in familial combined hyperlipidemia associated with Apolipoprotein B. Arterioscler Thromb Vasc Biol, 2002. 22(2): p. 283-8.
  74. Chait, A. and J.D. Brunzell, Severe hypertriglyceridemia:Role of familial and acquired disorders. Metabolism: Clinical and Experimental, 1983. 32: p. 209-214.
  75. Herink, M. and M.K. Ito, Medication Induced Changes in Lipid and Lipoproteins, in Endotext, K.R. Feingold, et al., Editors. 2018: South Dartmouth (MA).
  76. Feingold, K.R., The Effect of Inflammation and Infection on Lipids and Lipoproteins, in Endotext, K.R., Feingold, et al., Editors. 2022: South Dartmouth (MA).
  77. Feingold, K.R., Obesity and Dyslipidemia, in Endotext, K.R. Feingold, et al., Editors. 2023: South Dartmouth (MA).
  78. Feingold, K.R., The Effect of Endocrine Disorders on Lipids and Lipoproteins, in Endotext, K.R. Feingold, et al., Editors. 2023: South Dartmouth (MA).
  79. Chait, A., Secondary hyperlipidemia. Journal of Clinical Pathology, 1973. 26(suppl 5): p. 68-71.
  80. Chait, A. and J.D. Brunzell, Acquired hyperlipidemia (secondary dyslipoproteinemia). Endocrinology and Metabolism Clinics of North America, 1990. 19: p. 259-278.
  81. Brunzell, J.D. and E.L. Bierman, Chylomicronemia syndrome.Interaction of genetic and acquired hypertriglyceridemia. Medical Clinics of North America, 1982. 66: p. 455-468.
  82. Chait, A., H.T. Robertson, and J.D. Brunzell, Chylomicronemia syndrome in diabetes mellitus. Diabetes Care, 1981. 4: p. 343-348.
  83. Brahm, A.J. and R.A. Hegele, Chylomicronaemia-current diagnosis and future therapies. Nat Rev Endocrinol, 2015.
  84. Brunzell, J. and S. Deeb, Familial lipoprotein lipase deficiency, Apo CII deficiency, and hepatic lipase deficiency, in The Metabolic and Molecular Basis of Inherited Disease, C. Scriver, et al., Editors. 2001, McGraw-Hill Book Co.: New York. p. 2789-2816.
  85. Rahalkar, A.R., et al., Novel LPL mutations associated with lipoprotein lipase deficiency: two case reports and a literature review. Can J Physiol Pharmacol, 2009. 87(3): p. 151-60.
  86. Martin-Campos, J.M., et al., Molecular analysis of chylomicronemia in a clinical laboratory setting: diagnosis of 13 cases of lipoprotein lipase deficiency. Clin Chim Acta, 2014. 429: p. 61-8.
  87. Blom, D.J., et al., Characterizing familial chylomicronemia syndrome: Baseline data of the APPROACH study. J Clin Lipidol, 2018. 12(5): p. 1234-1243 e5.
  88. Peterson, J., et al., Structural and functional consequences of missense mutations in exon 5 of the lipoprotein lipase gene. J Lipid Res, 2002. 43(3): p. 398-406.
  89. Nickerson, D.A., et al., DNA sequence diversity in a 9.7-kb region of the human lipoprotein lipase gene. Nat Genet, 1998. 19(3): p. 233-40.
  90. Breckenridge, W.C., et al., Hypertriglyceridemia associated with deficiency of Apolipoprotein C-II. New England Journal of Medicine, 1978. 298: p. 1265.
  91. Rabacchi, C., et al., Spectrum of mutations of the LPL gene identified in Italy in patients with severe hypertriglyceridemia. Atherosclerosis, 2015. 241(1): p. 79-86.
  92. Surendran, R.P., et al., Mutations in LPL, APOC2, APOA5, GPIHBP1 and LMF1 in patients with severe hypertriglyceridaemia. J Intern Med, 2012. 272(2): p. 185-96.
  93. Kristensen, K.K., et al., A disordered acidic domain in GPIHBP1 harboring a sulfated tyrosine regulates lipoprotein lipase. Proc Natl Acad Sci U S A, 2018. 115(26): p. E6020-E6029.
  94. Calandra, S., et al., APOA5 and triglyceride metabolism, lesson from human APOA5 deficiency. Curr Opin Lipidol, 2006. 17(2): p. 122-7.
  95. Peterfy, M., Lipase maturation factor 1: a lipase chaperone involved in lipid metabolism. Biochim Biophys Acta, 2012. 1821(5): p. 790-4.
  96. Baass, A., et al., Familial chylomicronemia syndrome: an under-recognized cause of severe hypertriglyceridaemia. J Intern Med, 2020. 287(4): p. 340-348.
  97. Dron, J.S. and R.A. Hegele, The evolution of genetic-based risk scores for lipids and cardiovascular disease. Curr Opin Lipidol, 2019. 30(2): p. 71-81.
  98. Parsons, S.K., et al., Asparaginase-associated lipid abnormalities in children with acute lymphoblastic leukemia. Blood, 1997. 89(6): p. 1886-95.
  99. Tozuka, M., et al., Characterization of hypertriglyceridemia induced by L-asparaginase therapy for acute lymphoblastic leukemia and malignant lymphoma. Ann Clin Lab Sci, 1997. 27(5): p. 351-7.
  100. Yadav, D. and C.S. Pitchumoni, Issues in hyperlipidemic pancreatitis. J Clin Gastroenterol, 2003. 36(1): p. 54-62.
  101. Brunzell, J.D. and H.G. Schrott, The interaction of familial and secondary causes of hypertriglyceridemia:Role in pancreatitis. Transactions of the Association of American Physicians, 1973. 86: p. 245-254.
  102. Brunzell, J.D. and H.G. Schrott, The interaction of familial and secondary causes of hypertriglyceridemia: role in pancreatitis. J Clin Lipidol, 2012. 6(5): p. 409-12.
  103. Johansen, C.T., et al., An increased burden of common and rare lipid-associated risk alleles contributes to the phenotypic spectrum of hypertriglyceridemia. Arterioscler Thromb Vasc Biol, 2011. 31(8): p. 1916-26.
  104. Schneider, W.J., et al., Familial dysbetalipoproteinemia. Abnormal binding of mutant Apoprotein E to low density lipoprotein receptors of human fibroblasts and membranes from liver and adrenal of rats, rabbits, and cows. J Clin Invest, 1981. 68(4): p. 1075-85.
  105. Vedhachalam, C., et al., The C-terminal lipid-binding domain of Apolipoprotein E is a highly efficient mediator of ABCA1-dependent cholesterol efflux that promotes the assembly of high-density lipoproteins. Biochemistry, 2007. 46(10): p. 2583-93.
  106. Siest, G., et al., Apolipoprotein E: an important gene and protein to follow in laboratory medicine. Clin Chem, 1995. 41(8 Pt 1): p. 1068-86.
  107. Weisgraber, K.H., S.C. Rall, Jr., and R.W. Mahley, Human E Apoprotein heterogeneity. Cysteine-arginine interchanges in the amino acid sequence of the Apo-E isoforms. J Biol Chem, 1981. 256(17): p. 9077-83.
  108. Smelt, A.H. and F. de Beer, Apolipoprotein E and familial dysbetalipoproteinemia: clinical, biochemical, and genetic aspects. Semin Vasc Med, 2004. 4(3): p. 249-57.
  109. Dong, L.M., et al., The carboxyl terminus in Apolipoprotein E2 and the seven amino acid repeat in Apolipoprotein E-Leiden: role in receptor-binding activity. J Lipid Res, 1998. 39(6): p. 1173-80.
  110. Mahley, R.W., Y. Huang, and S.C. Rall, Jr., Pathogenesis of type III hyperlipoproteinemia (dysbetalipoproteinemia). Questions, quandaries, and paradoxes. J Lipid Res, 1999. 40(11): p. 1933-49.
  111. Koopal, C., A.D. Marais, and F.L. Visseren, Familial dysbetalipoproteinemia: an underdiagnosed lipid disorder. Curr Opin Endocrinol Diabetes Obes, 2017. 24(2): p. 133-139.
  112. Koopal, C., et al., Autosomal dominant familial dysbetalipoproteinemia: A pathophysiological framework and practical approach to diagnosis and therapy. J Clin Lipidol, 2017. 11(1): p. 12-23 e1.
  113. Cenarro, A., et al., The p.Leu167del Mutation in APOE Gene Causes Autosomal Dominant Hypercholesterolemia by Down-regulation of LDL Receptor Expression in Hepatocytes. J Clin Endocrinol Metab, 2016. 101(5): p. 2113-21.
  114. Bea, A.M., et al., Contribution of APOE Genetic Variants to Dyslipidemia. Arterioscler Thromb Vasc Biol, 2023. 43(6): p. 1066-1077.
  115. Paquette, M., S. Bernard, and A. Baass, Diagnosis of remnant hyperlipidaemia. Curr Opin Lipidol, 2022. 33(4): p. 227-230.
  116. Pallazola, V.A., et al., Modern prevalence of dysbetalipoproteinemia (Fredrickson-Levy-Lees type III hyperlipoproteinemia). Arch Med Sci, 2020. 16(5): p. 993-1003.
  117. Mahley, R.W. and S.C. Rall, Jr., Type III hyperlipoproteinemia (dysbetalipoproteinemia):The role of Apolipoprotein E in normal and abnormal lipoprotein metabolism, in The Metabolic Basis of Inherited Disease, C.R. Scriver, et al., Editors. 1989, McGraw-Hill: New York. p. 1195.
  118. Brummer, D., et al., Expression of type III hyperlipoproteinemia in patients homozygous for Apolipoprotein E-2 is modulated by lipoprotein lipase and postprandial hyperinsulinemia. J Mol Med (Berl), 1998. 76(5): p. 355-64.
  119. Feussner, G. and R. Ziegler, Expression of type III hyperlipoproteinaemia in a subject with secondary hypothyroidism bearing the Apolipoprotein E2/2 phenotype. J Intern Med, 1991. 230(2): p. 183-6.
  120. Breslow, J.L., et al., Studies of familial type III hyperlipoproteinemia using as a genetic marker the ApoE phenotype E2/2. J Lipid Res, 1982. 23(8): p. 1224-35.
  121. Chait, A., et al., Type-III Hyperlipoproteinaemia ("remnant removal disease"). Insight into the pathogenetic mechanism. Lancet, 1977. 1(8023): p. 1176-8.
  122. Chait, A., et al., Impaired very low density lipoprotein and triglyceride removal in broad beta disease: comparison with endogenous hypertriglyceridemia. Metabolism, 1978. 27(9): p. 1055-66.
  123. Rothschild, M., et al., Pathognomonic Palmar Crease Xanthomas of Apolipoprotein E2 Homozygosity-Familial Dysbetalipoproteinemia. JAMA Dermatol, 2016. 152(11): p. 1275-1276.
  124. Albers, J.J., G.R. Warnick, and W.R. Hazzard, Type III hyperlipoproteinemia: a comparative study of current diagnostic techniques. Clin Chim Acta, 1977. 75(2): p. 193-204.
  125. Blom, D.J., F.H. O'Neill, and A.D. Marais, Screening for dysbetalipoproteinemia by plasma cholesterol and Apolipoprotein B concentrations. Clin Chem, 2005. 51(5): p. 904-7.
  126. Paquette, M., et al., A simplified diagnosis algorithm for dysbetalipoproteinemia. J Clin Lipidol, 2020. 14(4): p. 431-437.
  127. Morganroth, J., R.I. Levy, and D.S. Fredrickson, The biochemical, clinical, and genetic features of type III hyperlipoproteinemia. Ann Intern Med, 1975. 82(2): p. 158-74.
  128. Havel, R.J. and J.P. Kane, Primary dysbetalipoproteinemia:Predominancy of a specific Apoprotein species in triglyceride-rich lipoproteins. Proceedings of the National Academy of Sciences of the USA, 1973. 70: p. 2015.
  129. Koopal, C., et al., Vascular risk factors, vascular disease, lipids and lipid targets in patients with familial dysbetalipoproteinemia: a European cross-sectional study. Atherosclerosis, 2015. 240(1): p. 90-7.
  130. Mahley, R. and S. Rall, Type III Hyperlipoproteinemia (Dysbetalipoproteinemia): The Role of Apolipoprotein E in Normal and Abnormal Lipoprotein Metabolism, in The Metabolic & Molecular Bases of Inherited Disease, C. Scriver, et al., Editors. 2001, McGraw-Hill: New York. p. 2835-2862.
  131. Koopal, C., et al., The relation between Apolipoprotein E (APOE) genotype and peripheral artery disease in patients at high risk for cardiovascular disease. Atherosclerosis, 2016. 246: p. 187-92.
  132. Paquette, M., S. Bernard, and A. Baass, Dysbetalipoproteinemia Is Associated With Increased Risk of Coronary and Peripheral Vascular Disease. J Clin Endocrinol Metab, 2022. 108(1): p. 184-190.
  133. Garg, A., Clinical review#: Lipodystrophies: genetic and acquired body fat disorders. J Clin Endocrinol Metab, 2011. 96(11): p. 3313-25.
  134. Akinci, B., M. Sahinoz, and E. Oral, Lipodystrophy Syndromes: Presentation and Treatment, in Endotext, K.R. Feingold, et al., Editors. 2000: South Dartmouth (MA).
  135. Simha, V. and A. Garg, Inherited lipodystrophies and hypertriglyceridemia. Curr Opin Lipidol, 2009. 20(4): p. 300-8.
  136. Lightbourne, M. and R.J. Brown, Genetics of Lipodystrophy. Endocrinol Metab Clin North Am, 2017. 46(2): p. 539-554.
  137. Lotta, L.A., et al., Integrative genomic analysis implicates limited peripheral adipose storage capacity in the pathogenesis of human insulin resistance. Nat Genet, 2017. 49(1): p. 17-26.
  138. Guillin-Amarelle, C., et al., Type 1 familial partial lipodystrophy: understanding the Kobberling syndrome. Endocrine, 2016. 54(2): p. 411-421.
  139. Agarwal, A.K. and A. Garg, A novel heterozygous mutation in peroxisome proliferator-activated receptor-gamma gene in a patient with familial partial lipodystrophy. J Clin Endocrinol Metab, 2002. 87(1): p. 408-11.
  140. Ajluni, N., et al., Spectrum of disease associated with partial lipodystrophy: lessons from a trial cohort. Clin Endocrinol (Oxf), 2017. 86(5): p. 698-707.
  141. Subramanyam, L., V. Simha, and A. Garg, Overlapping syndrome with familial partial lipodystrophy, Dunnigan variety and cardiomyopathy due to amino-terminal heterozygous missense lamin A/C mutations. Clin Genet, 2010. 78(1): p. 66-73.
  142. Hussain, I. and A. Garg, Lipodystrophy Syndromes. Endocrinol Metab Clin North Am, 2016. 45(4): p. 783-797.
  143. Hussain, I., N. Patni, and A. Garg, Lipodystrophies, dyslipidaemias and atherosclerotic cardiovascular disease. Pathology, 2019. 51(2): p. 202-212.
  144. Jacob, K.N., et al., Phenotypic heterogeneity in body fat distribution in patients with atypical Werner's syndrome due to heterozygous Arg133Leu lamin A/C mutation. J Clin Endocrinol Metab, 2005. 90(12): p. 6699-706.
  145. Garg, A., et al., Atypical progeroid syndrome due to heterozygous missense LMNA mutations. J Clin Endocrinol Metab, 2009. 94(12): p. 4971-83.
  146. Calvo, M. and E. Martinez, Update on metabolic issues in HIV patients. Curr Opin HIV AIDS, 2014. 9(4): p. 332-9.
  147. Castelli, W.P., The triglyceride issue: a view from Framingham. Am Heart J, 1986. 112(2): p. 432-7.
  148. Harchaoui, K.E., et al., Triglycerides and cardiovascular risk. Curr Cardiol Rev, 2009. 5(3): p. 216-22.
  149. Langsted, A., et al., Nonfasting cholesterol and triglycerides and association with risk of myocardial infarction and total mortality: the Copenhagen City Heart Study with 31 years of follow-up. J Intern Med, 2011. 270(1): p. 65-75.
  150. Nordestgaard, B.G. and A. Varbo, Triglycerides and cardiovascular disease. Lancet, 2014. 384(9943): p. 626-35.
  151. Hokanson, J.E. and M.A. Austin, Plasma triglyceride level is a risk factor for cardiovascular disease independent of high-density lipoprotein cholesterol level: a meta-analysis of population-based prospective studies. J Cardiovasc Risk, 1996. 3(2): p. 213-9.
  152. Zilversmit, D.B., Atherogenesis:A postprandial phenomenon. Circulation, 1979. 60: p. 473-485.
  153. Zilversmit, D.B., Atherogenic nature of triglycerides, postprandial lipidemia, and triglyceride-rich remnant lipoproteins. Clin Chem, 1995. 41(1): p. 153-8.
  154. Nordestgaard, B.G. and D.B. Zilversmit, Large lipoproteins are excluded from the arterial wall in diabetic cholesterol-fed rabbits. J Lipid Res, 1988. 29(11): p. 1491-500.
  155. Nordestgaard, B.G., S. Stender, and K. Kjeldsen, Reduced atherogenesis in cholesterol-fed diabetic rabbits. Giant lipoproteins do not enter the arterial wall. Arteriosclerosis, 1988. 8(4): p. 421-8.
  156. Williams, K.J. and I. Tabas, The response-to-retention hypothesis of early atherogenesis. Arteriosclerosis, Thrombosis, and Vascular Biology, 1995. 15: p. 551-561.
  157. Williams, K.J. and I. Tabas, The response-to-retention hypothesis of atherogenesis reinforced. Curr Opin Lipidol, 1998. 9(5): p. 471-4.
  158. Krauss, R.M., Dense low density lipoproteins and coronary artery disease. Am J Cardiol, 1995. 75(6): p. 53B-57B.
  159. Olin-Lewis, K., et al., ApoC-III content of ApoB-containing lipoproteins is associated with binding to the vascular proteoglycan biglycan. J Lipid Res, 2002. 43(11): p. 1969-77.
  160. Chait, A., et al., Susceptibility of small, dense low density lipoproteins to oxidative modification in subjects with the atherogenic lipoprotein phenotype, pattern B. American Journal of Medicine, 1993. 94: p. 350-356.
  161. Tribble, D.L., et al., Oxidative susceptibility of LDL density subfractions is related to their ubiquinol-10 and à-tocopherol content. Proceedings of the National Academy of Sciences of the USA, 1994. 91: p. 1183-1187.
  162. Cornier, M.A., et al., The metabolic syndrome. Endocr Rev, 2008. 29(7): p. 777-822.
  163. Rosenson, R.S., et al., Genetics and causality of triglyceride-rich lipoproteins in atherosclerotic cardiovascular disease. J Am Coll Cardiol, 2014. 64(23): p. 2525-40.
  164. Do, R., et al., Common variants associated with plasma triglycerides and risk for coronary artery disease. Nat Genet, 2013. 45(11): p. 1345-52.
  165. Waterworth, D.M., et al., Genetic variants influencing circulating lipid levels and risk of coronary artery disease. Arterioscler Thromb Vasc Biol, 2010. 30(11): p. 2264-76.
  166. Teslovich, T.M., et al., Biological, clinical and population relevance of 95 loci for blood lipids. Nature, 2010. 466(7307): p. 707-13.
  167. Rip, J., et al., Lipoprotein lipase S447X: a naturally occurring gain-of-function mutation. Arterioscler Thromb Vasc Biol, 2006. 26(6): p. 1236-45.
  168. Ference, B.A., et al., Association of Triglyceride-Lowering LPL Variants and LDL-C-Lowering LDLR Variants With Risk of Coronary Heart Disease. JAMA, 2019. 321(4): p. 364-373.
  169. Pimstone, S.N., et al., Mutations in the gene for lipoprotein lipase. A cause for low HDL cholesterol levels in individuals heterozygous for familial hypercholesterolemia. Arterioscler Thromb Vasc Biol, 1995. 15(10): p. 1704-12.
  170. Burkhardt, R., et al., Trib1 is a lipid- and myocardial infarction-associated gene that regulates hepatic lipogenesis and VLDL production in mice. J Clin Invest, 2010. 120(12): p. 4410-4.
  171. Douvris, A., et al., Functional analysis of the TRIB1 associated locus linked to plasma triglycerides and coronary artery disease. J Am Heart Assoc, 2014. 3(3): p. e000884.
  172. Tg, et al., Loss-of-function mutations in APOC3, triglycerides, and coronary disease. N Engl J Med, 2014. 371(1): p. 22-31.
  173. Jorgensen, A.B., et al., Loss-of-function mutations in APOC3 and risk of ischemic vascular disease. N Engl J Med, 2014. 371(1): p. 32-41.
  174. Soufi, M., et al., Mutation screening of the APOA5 gene in subjects with coronary artery disease. J Investig Med, 2012. 60(7): p. 1015-9.
  175. Do, R., et al., Exome sequencing identifies rare LDLR and APOA5 alleles conferring risk for myocardial infarction. Nature, 2015. 518(7537): p. 102-6.
  176. Dewey, F.E., et al., Inactivating Variants in ANGPTL4 and Risk of Coronary Artery Disease. N Engl J Med, 2016. 374(12): p. 1123-33.
  177. Myocardial Infarction, G., et al., Coding Variation in ANGPTL4, LPL, and SVEP1 and the Risk of Coronary Disease. N Engl J Med, 2016. 374(12): p. 1134-44.
  178. Nordestgaard, B.G., R. Wootton, and B. Lewis, Selective retention of VLDL, IDL, and LDL in the arterial intima of genetically hyperlipidemic rabbits in vivo. Molecular size as a determinant of fractional loss from the intima-inner media. Arterioscler Thromb Vasc Biol, 1995. 15(4): p. 534-42.
  179. Norata, G.D., et al., Post-prandial endothelial dysfunction in hypertriglyceridemic subjects: molecular mechanisms and gene expression studies. Atherosclerosis, 2007. 193(2): p. 321-7.
  180. Malloy, M.J. and J.P. Kane, A risk factor for atherosclerosis: triglyceride-rich lipoproteins. Adv Intern Med, 2001. 47: p. 111-36.
  181. Mamo, J.C., S.D. Proctor, and D. Smith, Retention of chylomicron remnants by arterial tissue; importance of an efficient clearance mechanism from plasma. Atherosclerosis, 1998. 141 Suppl 1: p. S63-9.
  182. Havel, R.J. and R.S. Gordon, Jr., Idiopathic hyperlipemia: metabolic studies in an affected family. J Clin Invest, 1960. 39: p. 1777-90.
  183. Benlian, P., et al., Premature atherosclerosis in patients with familial chylomicronemia caused by mutations in the lipoprotein lipase gene. N Engl J Med, 1996. 335(12): p. 848-54.
  184. Zafrir, B., et al., Clinical features and outcomes of severe, very severe, and extreme hypertriglyceridemia in a regional health service. J Clin Lipidol, 2018. 12(4): p. 928-936.
  185. Austin, M.A. and J.E. Hokanson, Epidemiology of triglycerides, small dense low-density lipoprotein, and lipoprotein(a) as risk factors for coronary heart disease. Med Clin North Am, 1994. 78(1): p. 99-115.
  186. Goldberg, I.J., R.H. Eckel, and R. McPherson, Triglycerides and heart disease: still a hypothesis? Arterioscler Thromb Vasc Biol, 2011. 31(8): p. 1716-25.
  187. Brown, W.V., et al., Severe hypertriglyceridemia. J Clin Lipidol, 2012. 6(5): p. 397-408.
  188. Goldberg, A.S. and R.A. Hegele, Severe hypertriglyceridemia in pregnancy. J Clin Endocrinol Metab, 2012. 97(8): p. 2589-96.
  189. Brunzell, J.D., Lipoprotein lipase deficiency and other causes of the chylomicronemia syndrome, in The Metabolic Basis of Inherited Disease, C.R. Scriver, et al., Editors. 1989, McGraw-Hill: New York. p. 1165-1180.
  190. Tremblay, K., et al., Etiology and risk of lactescent plasma and severe hypertriglyceridemia. J Clin Lipidol, 2011. 5(1): p. 37-44.
  191. Lloret Linares, C., et al., Acute pancreatitis in a cohort of 129 patients referred for severe hypertriglyceridemia. Pancreas, 2008. 37(1): p. 13-2.
  192. Cameron, J.L., et al., Acute pancreatitis with hyperlipidemia:The incidence of lipid abnormalities in acute pancreatitis. Annals of Surgery, 1973. 177: p. 483-489.
  193. Farmer, R.G., et al., Hyperlipoproteinemia and pancreatitis. American Journal of Medicine, 1973. 54: p. 161-165.
  194. Wang, Q., et al., Elevated Serum Triglycerides in the Prognostic Assessment of Acute Pancreatitis: A Systematic Review and Meta-Analysis of Observational Studies. J Clin Gastroenterol, 2017. 51(7): p. 586-593.
  195. Mirtallo, J.M., et al., State of the art review: Intravenous fat emulsions: Current applications, safety profile, and clinical implications. Ann Pharmacother, 2010. 44(4): p. 688-700.
  196. Devaud, J.C., et al., Hypertriglyceridemia: a potential side effect of propofol sedation in critical illness. Intensive Care Med, 2012. 38(12): p. 1990-8.
  197. Havel, R.J., Pathogenesis, differentiation and management of hypertriglyceridemia. Adv Intern Med, 1969. 15: p. 117-54.
  198. Yang, F., et al., The role of free fatty acids, pancreatic lipase and Ca+ signalling in injury of isolated acinar cells and pancreatitis model in lipoprotein lipase-deficient mice. Acta Physiol (Oxf), 2009. 195(1): p. 13-28.
  199. Tsuang, W., et al., Hypertriglyceridemic pancreatitis: presentation and management. Am J Gastroenterol, 2009. 104(4): p. 984-91.
  200. Valdivielso, P., A. Ramirez-Bueno, and N. Ewald, Current knowledge of hypertriglyceridemic pancreatitis. Eur J Intern Med, 2014. 25(8): p. 689-94.
  201. Saharia, P., et al., Acute pancreatitis with hyperlipemia: studies with an isolated perfused canine pancreas. Surgery, 1977. 82(1): p. 60-7.
  202. Seplowitz, A.H., S. Chien, and F.R. Smith, Effects of lipoproteins on plasma viscosity. Atherosclerosis, 1981. 38(1-2): p. 89-95.
  203. Chang, Y.T., et al., Association of cystic fibrosis transmembrane conductance regulator (CFTR) mutation/variant/haplotype and tumor necrosis factor (TNF) promoter polymorphism in hyperlipidemic pancreatitis. Clin Chem, 2008. 54(1): p. 131-8.
  204. Ivanova, R., et al., Triglyceride levels and Apolipoprotein E polymorphism in patients with acute pancreatitis. Hepatobiliary Pancreat Dis Int, 2012. 11(1): p. 96-101.
  205. Durrington, P., Dyslipidaemia. Lancet, 2003. 362(9385): p. 717-31.
  206. Parker, F., et al., Evidence for the chylomicron origin of lipids accumulating in diabetic eruptive xanthomas:A correlative lipid biochemical, histochemical and electron microscopic study. Journal of Clinical Investigation, 1970. 49: p. 2172-2187.
  207. Rosenson, R.S., et al., Hypertriglyceridemia and other factors associated with plasma viscosity. Am J Med, 2001. 110(6): p. 488-92.
  208. Inokuchi, R., et al., Hypertriglyceridemia as a possible cause of coma: a case report. J Med Case Rep, 2012. 6: p. 412.
  209. Feingold, K.R., Triglyceride Lowering Drugs, in Endotext, K.R. Feingold, et al., Editors. 2021: South Dartmouth (MA).
  210. Feingold K.R., The Effect of Diet on Cardiovascular Disease and Lipid and Lipoprotein Levels, in Endotext, K.R. Feingold, et al., Editors. 2021: South Dartmouth (MA).
  211. Scandinavian Simvastatin Survival Study Group, Randomized trial of cholesterol lowering in 4444 patients with coronary heart disease: The Scandinavian Simvastatin Survival Study (4S). Lancet, 1994. 344: p. 1383-1389.
  212. Cannon, C.P., et al., Ezetimibe Added to Statin Therapy after Acute Coronary Syndromes. N Engl J Med, 2015.
  213. Koskinen, P., et al., Coronary heart disease incidence in NIDDM patients in the Helsinki Heart Study. Diab Care, 1992. 15: p. 825-829.
  214. ACCORD Study Group, et al., Effects of combination lipid therapy in type 2 diabetes mellitus. N Engl J Med, 2010. 362(17): p. 1563-74.
  215. Keech, A., et al., Effects of long-term fenofibrate therapy on cardiovascular events in 9795 people with type 2 diabetes mellitus (the FIELD study): randomised controlled trial. Lancet, 2005. 366(9500): p. 1849-61.
  216. Bezafibrate Infarction Prevention, s., Secondary prevention by raising HDL cholesterol and reducing triglycerides in patients with coronary artery disease. Circulation, 2000. 102(1): p. 21-7.
  217. Camejo, G., Phase 2 clinical trials with K-877 (pemafibrate): A promising selective PPAR-alpha modulator for treatment of combined dyslipidemia. Atherosclerosis, 2017. 261: p. 163-164.
  218. Das Pradhan, A., et al., Triglyceride Lowering with Pemafibrate to Reduce Cardiovascular Risk. N Engl J Med, 2022. 387(21): p. 1923-1934.
  219. Aung, T., et al., Associations of Omega-3 Fatty Acid Supplement Use With Cardiovascular Disease Risks: Meta-analysis of 10 Trials Involving 77917 Individuals. JAMA Cardiol, 2018. 3(3): p. 225-234.
  220. Bhatt, D.L., et al., Cardiovascular Risk Reduction with Icosapent Ethyl for Hypertriglyceridemia. N Engl J Med, 2018.
  221. Nicholls, S.J., et al., Effect of High-Dose Omega-3 Fatty Acids vs Corn Oil on Major Adverse Cardiovascular Events in Patients at High Cardiovascular Risk: The STRENGTH Randomized Clinical Trial. JAMA, 2020. 324(22): p. 2268-2280.
  222. Kalstad, A.A., et al., Effects of n-3 Fatty Acid Supplements in Elderly Patients After Myocardial Infarction: A Randomized, Controlled Trial. Circulation, 2021. 143(6): p. 528-539.
  223. Goff, Z.D. and S.E. Nissen, N-3 polyunsaturated fatty acids for cardiovascular risk. Curr Opin Cardiol, 2022. 37(4): p. 356-363.
  224. Mason, R.P., S.C.R. Sherratt, and R.H. Eckel, Omega-3-fatty acids: Do they prevent cardiovascular disease? Best Pract Res Clin Endocrinol Metab, 2023. 37(3): p. 101681.
  225. Zambon, A., et al., New and Emerging Therapies for Dyslipidemia. Endocrinol Metab Clin North Am, 2022. 51(3): p. 635-653.
  226. Haller, J.F., et al., ANGPTL8 requires ANGPTL3 to inhibit lipoprotein lipase and plasma triglyceride clearance. J Lipid Res, 2017. 58(6): p. 1166-1173.
  227. Rosenson, R.S., et al., Evinacumab in severe hypertriglyceridemia with or without lipoprotein lipase pathway mutations: a phase 2 randomized trial. Nat Med, 2023. 29(3): p. 729-737.
  228. Chaudhry, R., A. Viljoen, and A.S. Wierzbicki, Pharmacological treatment options for severe hypertriglyceridemia and familial chylomicronemia syndrome. Expert Rev Clin Pharmacol, 2018. 11(6): p. 589-598.
  229. Capell, W.H. and R.H. Eckel, Treatment of hypertriglyceridemia. Curr Diab Rep, 2006. 6(3): p. 230-40.
  230. Yuan, G., K.Z. Al-Shali, and R.A. Hegele, Hypertriglyceridemia: its etiology, effects and treatment. CMAJ, 2007. 176(8): p. 1113-20.
  231. Whayne, T.F., Jr. and J.M. Felts, Activation of lipoprotein lipase. Effects of rat serum lipoprotein fractions and heparin. Circ Res, 1970. 27(6): p. 941-51.
  232. Weintraub, M., et al., Continuous intravenous heparin administration in humans causes a decrease in serum lipolytic activity and accumulation of chylomicrons in circulation. J Lipid Res, 1994. 35(2): p. 229-38.
  233. Whayne, T.F., Jr., Concerns about heparin therapy for hypertriglyceridemia. Arch Intern Med, 2010. 170(1): p. 108-9; author reply 109.
  234. Goldberg, I.J., Lipoprotein lipase and lipolysis: central roles in lipoprotein metabolism and atherogenesis. J Lipid Res, 1996. 37(4): p. 693-707.
  235. Aryal, M.R., et al., Acute pancreatitis owing to very high triglyceride levels treated with insulin and heparin infusion. BMJ Case Rep, 2013. 2013.
  236. Khan, A.S., S.U. Latif, and M.A. Eloubeidi, Controversies in the etiologies of acute pancreatitis. JOP, 2010. 11(6): p. 545-52.
  237. Coskun, A., et al., Treatment of hypertriglyceridemia-induced acute pancreatitis with insulin. Prz Gastroenterol, 2015. 10(1): p. 18-22.
  238. Mikhail, N., et al., Treatment of severe hypertriglyceridemia in nondiabetic patients with insulin. Am J Emerg Med, 2005. 23(3): p. 415-7.
  239. Jabbar, M.A., M.I. Zuhri-Yafi, and J. Larrea, Insulin therapy for a non-diabetic patient with severe hypertriglyceridemia. J Am Coll Nutr, 1998. 17(5): p. 458-61.
  240. Thuzar, M., et al., Extreme hypertriglyceridemia managed with insulin. J Clin Lipidol, 2014. 8(6): p. 630-4.
  241. Szczepiorkowski, Z.M., et al., Guidelines on the use of therapeutic apheresis in clinical practice--evidence-based approach from the Apheresis Applications Committee of the American Society for Apheresis. J Clin Apher, 2010. 25(3): p. 83-177.
  242. Stefanutti, C. and U. Julius, Treatment of primary hypertriglyceridemia states - General approach and the role of extracorporeal methods. Atheroscler Suppl, 2015. 18: p. 85-94.
  243. Furuya, T., et al., Plasma exchange for hypertriglyceridemic acute necrotizing pancreatitis: report of two cases. Ther Apher, 2002. 6(6): p. 454-8.
  244. Click, B., et al., The role of apheresis in hypertriglyceridemia-induced acute pancreatitis: A systematic review. Pancreatology, 2015. 15(4): p. 313-20.
  245. Webb, C.B., et al., Effect of TPE vs medical management on patient outcomes in the setting of hypertriglyceridemia-induced acute pancreatitis with severely elevated triglycerides. J Clin Apher, 2021. 36(5): p. 719-726.
  246. Koutroumpakis, E., et al., Management and outcomes of acute pancreatitis patients over the last decade: A US tertiary-center experience. Pancreatology, 2017. 17(1): p. 32-40.
  247. Huang, C., et al., Clinical features and treatment of hypertriglyceridemia-induced acute pancreatitis during pregnancy: A retrospective study. J Clin Apher, 2016. 31(6): p. 571-578.
  248. Wierzbicki, A.S., T.M. Reynolds, and M.A. Crook, Usefulness of Orlistat in the treatment of severe hypertriglyceridemia. Am J Cardiol, 2002. 89(2): p. 229-31.
  249. Tolentino, M.C., et al., Combination of gemfibrozil and orlistat for treatment of combined hyperlipidemia with predominant hypertriglyceridemia. Endocr Pract, 2002. 8(3): p. 208-12.
  250. Davidson, M., et al., The burden of familial chylomicronemia syndrome: interim results from the IN-FOCUS study. Expert Rev Cardiovasc Ther, 2017. 15(5): p. 415-423.
  251. Rouis, M., et al., Therapeutic response to medium-chain triglycerides and omega-3 fatty acids in a patient with the familial chylomicronemia syndrome. Arterioscler Thromb Vasc Biol, 1997. 17(7): p. 1400-6.
  252. Brunzell, J.D., Familial Lipoprotein Lipase Deficiency, in GeneReviews at GeneTests: Medical Genetics Information Resource. 2011, University of Washington: Seattle. p. 1997-2010.
  253. Brunzell JD, D.S., Familial lipoprotein lipase deficiency, Apo CII deficiency and hepatic lipase deficiency., in The Metabolic and Molecular Basis of Inherited Disease, 8th edition2001, McGraw-Hill Book Co.: New York. p. 2789-2816.
  254. Patni, N., C. Quittner, and A. Garg, Orlistat Therapy for Children With Type 1 Hyperlipoproteinemia: A Randomized Clinical Trial. J Clin Endocrinol Metab, 2018. 103(6): p. 2403-2407.
  255. Blackett, P., et al., Lipoprotein abnormalities in compound heterozygous lipoprotein lipase deficiency after treatment with a low-fat diet and orlistat. J Clin Lipidol, 2013. 7(2): p. 132-9.
  256. Tsai, E.C., et al., Potential of essential fatty acid deficiency with extremely low fat diet in lipoprotein lipase deficiency during pregnancy: A case report. BMC Pregnancy Childbirth, 2004. 4(1): p. 27.
  257. Al-Shali, K., et al., Successful pregnancy outcome in a patient with severe chylomicronemia due to compound heterozygosity for mutant lipoprotein lipase. Clin Biochem, 2002. 35(2): p. 125-30.
  258. Carpentier, A.C., et al., Effect of alipogene tiparvovec (AAV1-LPL(S447X)) on postprandial chylomicron metabolism in lipoprotein lipase-deficient patients. J Clin Endocrinol Metab, 2012. 97(5): p. 1635-44.
  259. Sanada, M., et al., Substitution of transdermal estradiol during oral estrogen-progestin therapy in postmenopausal women: effects on hypertriglyceridemia. Menopause, 2004. 11(3): p. 331-6.
  260. Hemelaar, M., et al., Oral, more than transdermal, estrogen therapy improves lipids and lipoprotein(a) in postmenopausal women: a randomized, placebo-controlled study. Menopause, 2003. 10(6): p. 550-8.
  261. Hsu, S.Y., et al., Laparoscopic bariatric surgery for the treatment of severe hypertriglyceridemia. Asian J Surg, 2015. 38(2): p. 96-101.
  262. Gaudet, D., et al., Antisense Inhibition of Apolipoprotein C-III in Patients with Hypertriglyceridemia. N Engl J Med, 2015. 373(5): p. 438-47.

 

Monogenic Disorders Altering HDL Levels

ABSTRACT

 

Very low HDL-C levels (<20mg/dL) may be due to severe elevations in triglycerides, very poorly controlled diabetes, inflammation, infections, malignancy, liver disease, and certain medications such as anabolic steroids. Additionally, variants in multiple genes that each have a small effect but cumulatively lead to a decrease in HDL-C can result in very low HDL-C levels. Finally, rare monogenic disorders such as familial hypoalphalipoproteinemia, Tangier disease, and lecithin acyltransferase (LCAT) deficiency can lead to very low HDL-C levels. In this chapter we discuss the lipid abnormalities and clinical features of these monogenic disorders causing very low HDL-C levels. An elevated concentration of apo A-I and apo A-II is called hyperalphalipoproteinemia (HALP). HALP is classified as moderate (HDL-C levels between 80 and 100 mg/dL) or severe (HDL-C levels > 100 mg/dL). HALP is a heterogeneous condition caused by a variety of genetic and secondary conditions (for example ethanol abuse, primary biliary cirrhosis, multiple lipomatosis, emphysema, exercise, and certain drugs such as estrogens). In many individuals HALP has a polygenic origin. Monogenic HALP includes CETP deficiency, hepatic lipase deficiency, endothelial lipase deficiency, and loss of function mutations in SRB1. In this chapter we discuss the lipid abnormalities and clinical features of these monogenic disorders causing HALP.

 

LOW HDL CONDITIONS

 

The inverse relationship between HDL-C and ASCVD risk is well established but it should be recognized that while this association is consistently observed recent genetic and cardiovascular outcome studies suggest that this association is not causal (1). However, as discussed below major reductions in HDL-C induced by specific monogenic disorders may increase the risk of ASCVD.

 

Isolated low HDL-C levels can occur; however, it is more commonly found in association with hypertriglyceridemia and/or elevated apo B levels, typically as part of the obesity/metabolic syndrome (2). Patients with very low HDL-C (<20 mg/dL) in the absence of severe hypertriglyceridemia, very poorly controlled diabetes, inflammation, infections, malignancy, liver disease, anabolic steroids, or a paradoxical response to PPAR agonists are very rare (<1% of the population) (3,4). These individuals may have a very rare monogenic disorder associated with marked HDL deficiency, including familial hypoalphalipoproteinemia, Tangier disease, and lecithin acyltransferase (LCAT) deficiency. Table 1 summarizes the genetic, lipid, and clinical features of these monogenic low HDL conditions. Inheritance is autosomal co-dominant with heterozygotes having decreases in HDL-C levels approximately midway between normal and homozygotes (3). In some individuals the decrease in HDL-C can be polygenic i.e., variants in multiple genes that each have a small effect but cumulatively lead to a decrease in HDL-C (5).

 

Table 1. Characteristics of Monogenic Low HDL Syndromes

 

Effected genes

Lipids

Clinical features

Familial hypoalpha-lipoproteinemia

apo A-I/apo C-III/ apo A-IV

apo A-I/apo C-III

apo A-I

Apo AI undetectable, marked deficiency in HDL-C, low – normal triglycerides, normal LDL-C

Xanthomas Premature ASCVD Corneal manifestations

Tangier disease

ABCA1

HDL species exclusively preß-1, HDL-C <5 mg/dL

LDL-C low (half normal)

Hepatosplenomegaly

Enlarged tonsils

Neuropathy

ASCVD (6-7th decade)

LCAT deficiency

LCAT

HDL-C <10 mg/dL

apo A-I 20-30 mg/dL

<36% cholesteryl esters

Low LDL-C

Presence of Lp-X particles

FLD develop corneal opacities (“fish eye”), normochromic anemia and proteinuric end stage renal disease

 

FED only develop corneal opacities

Inheritance is autosomal co-dominant with heterozygotes having decreases in HDL-C levels approximately midway between normal and homozygotes (3). FLD- Familial Lecithin: Cholesteryl Ester Acyltransferase Deficiency; FED- Fish Eye Disease

 

Familial Hypoalphalipoproteinemia  

 

Familial hypoalphalipoproteinemia is a heterogeneous group of apolipoprotein A-I (apo A-I) deficiency states. This disorder is the rarest cause of  monogenic severe HDL deficiency (6). These various conditions are characterized by the specific apolipoprotein genes that are affected on the apo A-I/C-III/A-IV gene cluster (3). The genes for these 3 apolipoproteins (apo A-I, apo C-III, and apo A-IV) are grouped together in a cluster on human chromosome 11. In patients with apo A-I/C-III/A-IV deficiency, apoA-1 is undetectable in the plasma and is associated with marked deficiency in HDL-C, low triglyceride levels (due to apo C-III deficiency), and normal LDL-C levels (3). Heterozygotes have plasma HDL-C, apo A-I, apo A-IV, and apo C-III levels that are about 50% of normal (3). This condition is associated with aggressive, premature ASCVD. Additionally, there is evidence of mild fat malabsorption due to deficiency of apo A-IV. Patients with apo A-I/C-III deficiency have undetectable apo A-I and a similar lipid profile as those with apo A-I/C-III/A-IV deficiency (3). This condition is also associated with premature ASCVD. It is distinguished from the former by presence of planar xanthomas and absence of fat malabsorption (since apo A-IV is present). Familial apo A-I deficiency is itself a heterogeneous group of disorders associated with numerous Apo A-I mutations (3). Common manifestations include undetectable plasma Apo A-I, marked HDL deficiency with normal LDL-C and triglyceride levels, xanthomas (planar, tendon, and/or tubero-eruptive depending on the specific gene mutation), and premature ASCVD. Some forms of the disease are also associated with corneal manifestations, including corneal arcus and corneal opacification. One of the interesting manifestations of familial apo A-I deficiency is that levels of apo A-IV and apo E containing HDL particles are only modestly reduced, with preserved electrophoretic mobility and particle size (7).

 

It is notable that familial hypoalphalipoproteinemia is associated with an increased risk of premature ASCVD presumably due to the marked deficiency in Apo A-I and HDL. Given the increased ASCVD risk associated with Apo A-I deficiency, treatment is directed towards aggressive reduction of LDL-C and non-HDL-C levels and reducing other cardiovascular risk factors.

 

Some mutations in Apo A-I are associated with low HDL-C levels and hereditary amyloidosis and are the second most frequent cause of familial amyloidosis (6,8). Note that HDL-C levels are not always decreased in patients with familial amyloidosis secondary to Apo A-I mutations. The N-terminal fragment of the mutated protein is found in the amyloid fragments.

 

Tangier Disease

 

Tangier disease is due to mutations in the gene that codes for ATP-Binding Cassette transporter A1 (ABCA1) and is inherited in an autosomal co-dominant manner (9,10).  Fredrickson first reported this condition in two patients who hailed from Tangier Island in the Chesapeake Bay, for which the disorder is named. ABCA1 facilitates efflux of intracellular cholesterol from peripheral cells to lipid poor A1, the key first step of reverse cholesterol transport (11). As such, this disorder is characterized by severe deficiency of HDL-C (HDL-C <5 mg/dL) and the presence of only thepreß-1 HDL fraction of HDL (10). The poorly lipidated Apo A-I is rapidly catabolized by the kidney. These patients also demonstrate moderate hypertriglyceridemia and low LDL-C levels (10). The decrease in LDL-C is likely due to absence of the transfer of cholesterol from HDL to LDL. Studies have also suggested that an increase in LDL uptake by the liver also occurs (12). The increase in triglycerides may be due to the failure of HDL to provide co-factors that increase lipoprotein lipase activity. Additionally, ABCA1 deficiency in the liver increases triglyceride secretion and hepatic angiopoietin-like protein 3 secretion which could inhibit lipoprotein lipase activity leading to an increase in triglycerides (12,13).

 

Since ABCA1 deficiency impairs free cholesterol efflux from cells, there is accumulation of cholesterol esters in many tissues throughout the body (10). Classically, patients present with hepatosplenomegaly and enlarged yellow-orange hyperplastic tonsils, however, a wide spectrum of phenotypic manifestations is now appreciated with considerable variability in terms of clinical severity and organ involvement (9,10). Peripheral neuropathies are also a common complication and may be relapsing-remitting or chronic progressive (9,10). Tangier disease patients appear to have an increased risk of premature ASCVD, though not as pronounced as those with familial hypoalphalipoproteinemia (3,9,14). When the non-HDL-C levels are greater than 70mg/dL patients with Tangier disease are at higher risk of ASCVD whereas when the non-HDL-C levels are less than 70mg/dL ASCVD is low (9). Less common complications include corneal opacities and hematological manifestations such as thrombocytopenia and hemolytic anemia (9,10).

 

Individuals who are heterozygous for ABCA1 mutations have HDL-C levels that are variable but approximately 50% of normal with normal levels of preß-1 HDL but a deficiency of large α-1 and α-2 HDL particles (10). Cholesterol efflux capacity in heterozygotes has been reported as ~50% of normal. A mutation in one ABCA1 allele has been associated with increased risk of ASCVD in some studies and with no increase in ASCVD risk in other studies (15-20). Different mutations in ABCA1 result in varying HDL-C levels and phenotypes, which might explain the difference in ASCVD risk (21).

 

While Tangier patients manifest characteristically low HDL-C and Apo A-I, this lipid/lipoprotein phenotype is not adequate to make the diagnosis. ABCA1 gene sequence analysis is the preferred test to make the diagnosis of Tangier disease (10). Alternatively, non-denaturing two-dimensional electrophoresis followed by anti-apo A-I immunoblotting demonstrates only preβ1-HDL.

 

Currently, there is no specific treatment for Tangier disease (10). In fact, HDL-C raising therapies such as niacin and fibrates have proven ineffective in patients with this condition (22). Even HDL infusion was not beneficial (23). The major clinical issue in Tangier patients is disabling neuropathy; however, there is no effective intervention to manage this complication (10). Aggressive LDL-C lowering and treatment of other risk factors for atherosclerosis is recommended (10).

 

LCAT Deficiency  

 

LCAT is an enzyme that is bound primarily to HDL, with some also found on LDL (24,25). It facilitates cholesterol esterification by transferring a fatty acid from phosphatidyl choline to cholesterol (24,25). The hydrophobic cholesteryl esters are then sequestered in the core of the lipoprotein particles. LCAT is critical in the maturation of HDL particles. LCAT deficiency is an autosomal co-dominant disorder that manifests as either familial LCAT deficiency (FLD) or fish-eye disease in homozygotes (FED) (24,25). In FLD, mutations in LCAT lead to the inability of LCAT to esterify cholesterol in both HDL and LDL, whereas in FED, mutations in LCAT lead to the inability of LCAT to esterify cholesterol in HDL but the ability of LCAT to esterify cholesterol in LDL is preserved (24,25). Patients with FLD have virtually no cholesterol esters in the circulation while patients with FED have subnormal levels of cholesterol esters carried in apo B containing lipoproteins (24,25). Heterozygotes having decreases in HDL-C levels approximately midway between normal and homozygotes.  

 

Individuals with FLD develop corneal opacities (“fish eye”), normochromic hemolytic anemia (due to cholesterol enrichment of red blood cell membranes), mild thrombocytopenia, and proteinuric end stage renal disease, which is the major cause of morbidity and mortality (24,25). The corneal opacities begin early in life and some patients may need corneal transplants. The rate of development of renal disease is variable but in a large cohort renal failure occurred at a median age of 46 years (26). Patients with FED generally only manifest corneal opacities (24,25).

 

The lipid and lipoprotein profile in patients with FLD usually demonstrates low HDL-C levels (frequently <10 mg/dL) (24,27). In one cohort patients with FED tend to have higher HDL-C levels but in a large systematic review HDL-C levels were similar in patients with FLD and FED (24,27). LDL-C levels tend to be low in FLD and FED while triglyceride levels are increased (24,27). Lipoprotein X (Lp-X) particles are present in patients with FLD but not in patients with FED (24). Lp-X is a multilamellar vesicle with an aqueous core. It is primarily composed of free cholesterol and phospholipid with very little protein (albumin in the core and apolipoprotein C on the surface) and cholesteryl ester.

 

Given the association of Lp-X and kidney disease only with FLD (and not FED) and animal studies demonstrating the nephrotoxicity of Lp-X, it is likely that increased levels of Lp-X results in renal dysfunction in patients with FLD (25,28). Lp-X particles accumulate in the mesangial cells in the glomerulus and are thought to induce inflammation and breakdown of the basement membrane leading to proteinuria. It is notable that after renal transplantation in patients with FLD there is recurrence of renal damage in the transplanted kidney (26).

 

It is unclear as to whether LCAT deficiency is associated with an increased risk of ASCVD (25,29). Atherosclerosis imaging studies have yielded divergent data and the number of patients with FLD or FED studied is limited (25,29). In one study carriers of FLD mutations (i.e., heterozygotes) had a decrease in ASCVD while carriers of FED mutations had an increase in ASCVD (30). This may have been due to higher LDL-C levels in the carriers of FED mutations (30).

 

Current management of FLD focuses on managing the renal dysfunction. The associated kidney disease is traditionally managed with angiotensin-converting enzyme inhibitors, angiotensin receptor blockers, and a low-fat diet (25). Whether lipid lowering drugs are beneficial is unknown. Possible future therapies include enzyme replacement therapy with recombinant human LCAT, liver-directed LCAT gene therapy, small peptide or molecule activators of LCAT, and HDL mimetics (31,32). Infusions of recombinant human LCAT has improved the anemia and most parameters of renal function in a patient with FLD (33). Administration of CER-001, an apolipoprotein A1 (apoA-1)–containing HDL mimetic, has been shown to have beneficial effects on kidney and eye disease in a patient with LCAT deficiency (34).  

 

Approach to the Patient with Low HDL-C Levels

 

When encountering a patient with very low HDL-C levels it is important to first determine if this is a new abnormality or has been present for a long time. If prior HDL-C levels are normal, this excludes a primary monogenic etiology. If the decrease in HDL-C is new, one should consider the possibility of very poorly controlled diabetes, inflammation, infections, malignancy, liver disease, paraproteinemia, anabolic steroids, or a paradoxical response to PPAR agonists. Marked hypertriglyceridemia can also lead to very low HDL-C levels. 

 

In a patient with long-standing very low HDL-C levels without an identifiable secondary cause, one should consider a monogenic etiology. To evaluate potential monogenic causes, a detailed family history, with attention to HDL-C levels, is important. Obtaining lipid levels from relatives is very helpful. A focused physical examination, with particular attention to the skin, eyes, tonsils, and spleen may point to a specific monogenic disorder. Plasma apo A-I levels should be obtained. Individuals with familial hypoalphalipoproteinemia deficiency have undetectable plasma apo A-I. Patients with Tangier disease demonstrate very low apo A-I levels (<5 mg/dL). LCAT deficiency is associated with apo A-I levels that are low but substantially higher than the other monogenic etiologies. Patients with LCAT deficiency also have a higher ratio of free: total cholesterol in plasma and measurement of plasma free (unesterified) cholesterol can be helpful. Two-dimensional gel electrophoresis of plasma followed by immunoblotting with antibodies specific for apo A-I separates lipid-poor preß-HDL from lipid-rich–HDL and can be helpful in differentiating these disorders. Genetic analysis is indicated when a monogenic disorder is suspected.

 

HIGH HDL-C CONDITIONS (HYPERALPHALIPOPROTEINEMIA)

 

An elevated concentration of apo A-I and apo A-II is called hyperalphalipoproteinemia (HALP). HALP is classified as moderate (HDL-C levels between 80 and 100 mg/dL) or severe (HDL-C levels > 100 mg/dL). While it is well recognized that high HDL-C levels are associated with a decrease in ASCVD it should be noted that very high HDL-C levels are paradoxically associated with an increase in ASCVD (35,36).

 

HALP is a heterogeneous condition caused by a variety of genetic and secondary conditions (for example ethanol abuse, primary biliary cirrhosis, multiple lipomatosis, emphysema, exercise, and certain drugs such as estrogens). In many individuals, the very high HDL-C levels have a polygenic origin (5,37). Given the focus of this chapter, monogenic causes of HALP will be reviewed. Monogenic HALP includes CETP deficiency, hepatic lipase deficiency, endothelial lipase deficiency, and loss of function mutations in SRB1. Despite epidemiology that demonstrates an inverse relationship between HDL-C and ASCVD risk, some forms of familial HALP are paradoxically associated with increased cardiovascular risk.

 

HALP is generally identified incidentally after routine assessment of a lipid profile as it is not usually associated with any signs or symptoms. Generally, patients are asymptomatic and no medical therapy is required.

 

Cholesterol Ester Transfer Protein (CETP) Deficiency

 

CETP transfers cholesteryl esters from HDL particles to triglyceride rich lipoproteins and LDL in exchange for triglycerides (11). Individuals who are homozygous for CETP variants have very high HDL-C levels (>100mg/dL) while heterozygotes have moderately increased HDL-C levels (38-41). LDL-C and apo B levels may be normal or modestly decreased. The increase in HDL cholesterol are largely due to the accumulation of cholesterol esters (39). The decrease in LDL-C is due to the failure of cholesterol ester transport from HDL to apo B containing lipoproteins. There is a predominance of small LDL particles. Individuals who are heterozygotes for CETP mutations show modestly elevated HDL-C levels (38,39). In Japanese individuals with HDL-C levels > 100mg/dL 67% were demonstrated to have CETP gene mutations (42). CETP deficiency is the most important and frequent cause of HALP in Japan. CETP deficiency is common in other Asian populations but is relatively rare in other ethnic groups (39). Despite extensive studies the effect of CETP variants on the risk of ASCVD is uncertain (38-40,43). A variety of studies have indicated that a decrease in LDL-C and non-HDL-C levels (i.e. pro-atherogenic lipoproteins) rather than an increase in HDL-C induced by CETP variants underlies a potential beneficial effect on ASCVD (44).  

 

Endothelial Lipase (EL) Deficiency

 

Endothelial lipase (EL) is encoded by the LIPG gene and hydrolyzes phospholipids on HDL resulting in smaller HDL particles that are more rapidly metabolized (11). Genetic variants in LIPG have been identified Iin individuals with elevated HDL-C levels (38,39). As one would predict large HDL particles enriched in phospholipids are observed in individuals deficient in EL (39). Whether variants in LIPG leading to decreased EL activity and increased HDL-C levels reduces ASCVD risk is uncertain (38-40).

 

Hepatic Lipase (HL) Deficiency

 

Hepatic lipase (HL) is encoded by the LIPC gene and mediates the hydrolysis of triglycerides and phospholipids in intermediate density lipoproteins (IDL) and LDL leading to smaller particles (IDL is converted to LDL; LDL is converted from large LDL to small LDL) (11). It also mediates the hydrolysis of triglycerides and phospholipids in HDL resulting in smaller HDL particles (11). Several case reports of families with elevated HDL-C levels (HALP) caused by a genetically defined HL deficiency have been described (39,40). HL deficiency may also be associated with elevated triglycerides and cholesterol with increased intermediate density lipoproteins (IDL) (40,45). Several HL deficient individuals had premature ASCVD likely due to increased levels of apo B containing lipoproteins (40,45). Heterozygotes do not appear to have discrete lipoprotein abnormalities (45).

 

Scavenger Receptor Class B Type I (SR-BI)

 

Scavenger receptor class B type I (SR-BI) is encoded by the SCARB1 gene and facilitates the selective uptake of the cholesterol esters from HDL into the liver, adrenal, ovary, and testes (11). In macrophages and other cells, SR-B1 facilitates the efflux of cholesterol from the cell to HDL particles (11). SR-B1 deficient mice have an increase in atherosclerosis despite elevated HDL-C levels (46). Mutations in SCARB1 associated with decreased SR-B1 have been observed in individuals with high HDL-C levels (47-49). Heterozygotes have intermediate elevations of HDL-C between wild-type and homozygous individuals. Studies have suggested that some but not all mutations in SCARB1 result in an increased risk of ASCVD despite increased HDL-C levels (40,49). A decrease in adrenal function has been reported in some individuals with SCARB1 mutations likely due to a reduced ability of SR-B1 to facilitate cholesterol uptake into the adrenal glands (48,50). Abnormalities in platelet function have also been observed in some patients (50).

 

ACKNOWLEDGEMENTS

 

This work was supported by grants from the Northern California Institute for Research and Education.

 

REFERENCES

 

  1. Kjeldsen EW, Thomassen JQ, Frikke-Schmidt R. HDL cholesterol concentrations and risk of atherosclerotic cardiovascular disease - Insights from randomized clinical trials and human genetics. Biochim Biophys Acta Mol Cell Biol Lipids 2021; 1867:159063
  2. Feingold KR. Obesity and Dyslipidemia. In: Feingold KR, Anawalt B, Boyce A, Chrousos G, de Herder WW, Dhatariya K, Dungan K, Grossman A, Hershman JM, Hofland J, Kalra S, Kaltsas G, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, McGee EA, McLachlan R, Morley JE, New M, Purnell J, Sahay R, Singer F, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2023.
  3. Schaefer EJ, Anthanont P, Diffenderfer MR, Polisecki E, Asztalos BF. Diagnosis and treatment of high density lipoprotein deficiency. Prog Cardiovasc Dis 2016; 59:97-106
  4. Geller AS, Polisecki EY, Diffenderfer MR, Asztalos BF, Karathanasis SK, Hegele RA, Schaefer EJ. Genetic and secondary causes of severe HDL deficiency and cardiovascular disease. J Lipid Res 2018; 59:2421-2435
  5. Dron JS, Wang J, Low-Kam C, Khetarpal SA, Robinson JF, McIntyre AD, Ban MR, Cao H, Rhainds D, Dube MP, Rader DJ, Lettre G, Tardif JC, Hegele RA. Polygenic determinants in extremes of high-density lipoprotein cholesterol. J Lipid Res 2017; 58:2162-2170
  6. Zanoni P, von Eckardstein A. Inborn errors of apolipoprotein A-I metabolism: implications for disease, research and development. Curr Opin Lipidol 2020; 31:62-70
  7. Santos RD, Schaefer EJ, Asztalos BF, Polisecki E, Wang J, Hegele RA, Martinez LR, Miname MH, Rochitte CE, Da Luz PL, Maranhao RC. Characterization of high density lipoprotein particles in familial apolipoprotein A-I deficiency. J Lipid Res 2008; 49:349-357
  8. Joy T, Wang J, Hahn A, Hegele RA. APOA1 related amyloidosis: a case report and literature review. Clin Biochem 2003; 36:641-645
  9. Hooper AJ, Hegele RA, Burnett JR. Tangier disease: update for 2020. Curr Opin Lipidol 2020; 31:80-84
  10. Burnett JR, Hooper AJ, McCormick SPA, Hegele RA. Tangier Disease. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Mirzaa G, Amemiya A, eds. GeneReviews((R)). Seattle (WA) 2019.
  11. Feingold KR. Introduction to Lipids and Lipoproteins. In: Feingold KR, Anawalt B, Boyce A, Chrousos G, de Herder WW, Dhatariya K, Dungan K, Grossman A, Hershman JM, Hofland J, Kalra S, Kaltsas G, Koch C, Kopp P, Korbonits M, Kovacs CS, Kuohung W, Laferrere B, McGee EA, McLachlan R, Morley JE, New M, Purnell J, Sahay R, Singer F, Stratakis CA, Trence DL, Wilson DP, eds. Endotext. South Dartmouth (MA) 2021.
  12. Liu M, Chung S, Shelness GS, Parks JS. Hepatic ABCA1 and VLDL triglyceride production. Biochim Biophys Acta 2012; 1821:770-777
  13. Bi X, Pashos EE, Cuchel M, Lyssenko NN, Hernandez M, Picataggi A, McParland J, Yang W, Liu Y, Yan R, Yu C, DerOhannessian SL, Phillips MC, Morrisey EE, Duncan SA, Rader DJ. ATP-Binding Cassette Transporter A1 Deficiency in Human Induced Pluripotent Stem Cell-Derived Hepatocytes Abrogates HDL Biogenesis and Enhances Triglyceride Secretion. EBioMedicine 2017; 18:139-145
  14. Koseki M, Yamashita S, Ogura M, Ishigaki Y, Ono K, Tsukamoto K, Hori M, Matsuki K, Yokoyama S, Harada-Shiba M. Current Diagnosis and Management of Tangier Disease. J Atheroscler Thromb 2021; 28:802-810
  15. Bochem AE, van Wijk DF, Holleboom AG, Duivenvoorden R, Motazacker MM, Dallinga-Thie GM, de Groot E, Kastelein JJ, Nederveen AJ, Hovingh GK, Stroes ES. ABCA1 mutation carriers with low high-density lipoprotein cholesterol are characterized by a larger atherosclerotic burden. Eur Heart J 2013; 34:286-291
  16. van Dam MJ, de Groot E, Clee SM, Hovingh GK, Roelants R, Brooks-Wilson A, Zwinderman AH, Smit AJ, Smelt AH, Groen AK, Hayden MR, Kastelein JJ. Association between increased arterial-wall thickness and impairment in ABCA1-driven cholesterol efflux: an observational study. Lancet 2002; 359:37-42
  17. Schaefer EJ, Zech LA, Schwartz DE, Brewer HB, Jr. Coronary heart disease prevalence and other clinical features in familial high-density lipoprotein deficiency (Tangier disease). Ann Intern Med 1980; 93:261-266
  18. Frikke-Schmidt R, Nordestgaard BG, Stene MC, Sethi AA, Remaley AT, Schnohr P, Grande P, Tybjaerg-Hansen A. Association of loss-of-function mutations in the ABCA1 gene with high-density lipoprotein cholesterol levels and risk of ischemic heart disease. JAMA 2008; 299:2524-2532
  19. Frikke-Schmidt R. Genetic variation in the ABCA1 gene, HDL cholesterol, and risk of ischemic heart disease in the general population. Atherosclerosis 2010; 208:305-316
  20. Iatan I, Alrasadi K, Ruel I, Alwaili K, Genest J. Effect of ABCA1 mutations on risk for myocardial infarction. Curr Atheroscler Rep 2008; 10:413-426
  21. Brunham LR, Singaraja RR, Hayden MR. Variations on a gene: rare and common variants in ABCA1 and their impact on HDL cholesterol levels and atherosclerosis. Annu Rev Nutr 2006; 26:105-129
  22. Serfaty-Lacrosniere C, Civeira F, Lanzberg A, Isaia P, Berg J, Janus ED, Smith MP, Jr., Pritchard PH, Frohlich J, Lees RS, et al. Homozygous Tangier disease and cardiovascular disease. Atherosclerosis 1994; 107:85-98
  23. Schaefer EJ, Anderson DW, Zech LA, Lindgren FT, Bronzert TB, Rubalcaba EA, Brewer HB, Jr. Metabolism of high density lipoprotein subfractions and constituents in Tangier disease following the infusion of high density lipoproteins. J Lipid Res 1981; 22:217-228
  24. Pavanello C, Calabresi L. Genetic, biochemical, and clinical features of LCAT deficiency: update for 2020. Curr Opin Lipidol 2020; 31:232-237
  25. Saeedi R, Li M, Frohlich J. A review on lecithin:cholesterol acyltransferase deficiency. Clin Biochem 2015; 48:472-475
  26. Pavanello C, Ossoli A, Arca M, D'Erasmo L, Boscutti G, Gesualdo L, Lucchi T, Sampietro T, Veglia F, Calabresi L. Progression of chronic kidney disease in familial LCAT deficiency: a follow-up of the Italian cohort. J Lipid Res2020; 61:1784-1788
  27. Mehta R, Elias-Lopez D, Martagon AJ, Perez-Mendez OA, Sanchez MLO, Segura Y, Tusie MT, Aguilar-Salinas CA. LCAT deficiency: a systematic review with the clinical and genetic description of Mexican kindred. Lipids Health Dis 2021; 20:70
  28. Ossoli A, Neufeld EB, Thacker SG, Vaisman B, Pryor M, Freeman LA, Brantner CA, Baranova I, Francone NO, Demosky SJ, Jr., Vitali C, Locatelli M, Abbate M, Zoja C, Franceschini G, Calabresi L, Remaley AT. Lipoprotein X Causes Renal Disease in LCAT Deficiency. PLoS One 2016; 11:e0150083
  29. Norum KR, Remaley AT, Miettinen HE, Strom EH, Balbo BEP, Sampaio C, Wiig I, Kuivenhoven JA, Calabresi L, Tesmer JJ, Zhou M, Ng DS, Skeie B, Karathanasis SK, Manthei KA, Retterstol K. Lecithin:cholesterol acyltransferase: symposium on 50 years of biomedical research from its discovery to latest findings. J Lipid Res2020; 61:1142-1149
  30. Oldoni F, Baldassarre D, Castelnuovo S, Ossoli A, Amato M, van Capelleveen J, Hovingh GK, De Groot E, Bochem A, Simonelli S, Barbieri S, Veglia F, Franceschini G, Kuivenhoven JA, Holleboom AG, Calabresi L. Complete and Partial Lecithin:Cholesterol Acyltransferase Deficiency Is Differentially Associated With Atherosclerosis. Circulation 2018; 138:1000-1007
  31. Freeman LA, Karathanasis SK, Remaley AT. Novel lecithin: cholesterol acyltransferase-based therapeutic approaches. Curr Opin Lipidol 2020; 31:71-79
  32. Vitali C, Rader DJ, Cuchel M. Novel therapeutic opportunities for familial lecithin:cholesterol acyltransferase deficiency: promises and challenges. Curr Opin Lipidol 2023; 34:35-43
  33. Shamburek RD, Bakker-Arkema R, Auerbach BJ, Krause BR, Homan R, Amar MJ, Freeman LA, Remaley AT. Familial lecithin:cholesterol acyltransferase deficiency: First-in-human treatment with enzyme replacement. J Clin Lipidol 2016; 10:356-367
  34. Faguer S, Colombat M, Chauveau D, Bernadet-Monrozies P, Beq A, Delas A, Soler V, Labadens I, Huart A, Benlian P, Schanstra JP. Administration of the High-Density Lipoprotein Mimetic CER-001 for Inherited Lecithin-Cholesterol Acyltransferase Deficiency. Ann Intern Med 2021; 174:1022-1025
  35. Hirata A, Sugiyama D, Watanabe M, Tamakoshi A, Iso H, Kotani K, Kiyama M, Yamada M, Ishikawa S, Murakami Y, Miura K, Ueshima H, Okamura T, Evidence for Cardiovascular Prevention from Observational Cohorts in Japan Research G. Association of extremely high levels of high-density lipoprotein cholesterol with cardiovascular mortality in a pooled analysis of 9 cohort studies including 43,407 individuals: The EPOCH-JAPAN study. J Clin Lipidol 2018; 12:674-684 e675
  36. Madsen CM, Varbo A, Nordestgaard BG. Extreme high high-density lipoprotein cholesterol is paradoxically associated with high mortality in men and women: two prospective cohort studies. Eur Heart J 2017; 38:2478-2486
  37. Motazacker MM, Peter J, Treskes M, Shoulders CC, Kuivenhoven JA, Hovingh GK. Evidence of a polygenic origin of extreme high-density lipoprotein cholesterol levels. Arterioscler Thromb Vasc Biol 2013; 33:1521-1528
  38. Larach DB, Cuchel M, Rader DJ. Monogenic causes of elevated HDL cholesterol and implications for development of new therapeutics. Clin Lipidol 2013; 8:635-648
  39. Giammanco A, Noto D, Barbagallo CM, Nardi E, Caldarella R, Ciaccio M, Averna MR, Cefalu AB. Hyperalphalipoproteinemia and Beyond: The Role of HDL in Cardiovascular Diseases. Life (Basel) 2021; 11
  40. Kardassis D, Thymiakou E, Chroni A. Genetics and regulation of HDL metabolism. Biochim Biophys Acta Mol Cell Biol Lipids 2021; 1867:159060
  41. Yamashita S, Maruyama T, Hirano K, Sakai N, Nakajima N, Matsuzawa Y. Molecular mechanisms, lipoprotein abnormalities and atherogenicity of hyperalphalipoproteinemia. Atherosclerosis 2000; 152:271-285
  42. Hirano K, Yamashita S, Kuga Y, Sakai N, Nozaki S, Kihara S, Arai T, Yanagi K, Takami S, Menju M, et al. Atherosclerotic disease in marked hyperalphalipoproteinemia. Combined reduction of cholesteryl ester transfer protein and hepatic triglyceride lipase. Arterioscler Thromb Vasc Biol 1995; 15:1849-1856
  43. Yamashita S, Matsuzawa Y. Re-evaluation of cholesteryl ester transfer protein function in atherosclerosis based upon genetics and pharmacological manipulation. Curr Opin Lipidol 2016; 27:459-472
  44. Nicholls SJ, Ray KK, Nelson AJ, Kastelein JJP. Can we revive CETP-inhibitors for the prevention of cardiovascular disease? Curr Opin Lipidol 2022; 33:319-325
  45. Weissglas-Volkov D, Pajukanta P. Genetic causes of high and low serum HDL-cholesterol. J Lipid Res 2010; 51:2032-2057
  46. Trigatti B, Rayburn H, Vinals M, Braun A, Miettinen H, Penman M, Hertz M, Schrenzel M, Amigo L, Rigotti A, Krieger M. Influence of the high density lipoprotein receptor SR-BI on reproductive and cardiovascular pathophysiology. Proc Natl Acad Sci U S A 1999; 96:9322-9327
  47. Brunham LR, Tietjen I, Bochem AE, Singaraja RR, Franchini PL, Radomski C, Mattice M, Legendre A, Hovingh GK, Kastelein JJ, Hayden MR. Novel mutations in scavenger receptor BI associated with high HDL cholesterol in humans. Clin Genet 2011; 79:575-581
  48. Vergeer M, Korporaal SJ, Franssen R, Meurs I, Out R, Hovingh GK, Hoekstra M, Sierts JA, Dallinga-Thie GM, Motazacker MM, Holleboom AG, Van Berkel TJ, Kastelein JJ, Van Eck M, Kuivenhoven JA. Genetic variant of the scavenger receptor BI in humans. N Engl J Med 2011; 364:136-145
  49. Zanoni P, Khetarpal SA, Larach DB, Hancock-Cerutti WF, Millar JS, Cuchel M, DerOhannessian S, Kontush A, Surendran P, Saleheen D, Trompet S, Jukema JW, De Craen A, Deloukas P, Sattar N, Ford I, Packard C, Majumder A, Alam DS, Di Angelantonio E, Abecasis G, Chowdhury R, Erdmann J, Nordestgaard BG, Nielsen SF, Tybjaerg-Hansen A, Schmidt RF, Kuulasmaa K, Liu DJ, Perola M, Blankenberg S, Salomaa V, Mannisto S, Amouyel P, Arveiler D, Ferrieres J, Muller-Nurasyid M, Ferrario M, Kee F, Willer CJ, Samani N, Schunkert H, Butterworth AS, Howson JM, Peloso GM, Stitziel NO, Danesh J, Kathiresan S, Rader DJ, Consortium CHDE, Consortium CAE, Global Lipids Genetics C. Rare variant in scavenger receptor BI raises HDL cholesterol and increases risk of coronary heart disease. Science 2016; 351:1166-1171
  50. Chadwick AC, Sahoo D. Functional genomics of the human high-density lipoprotein receptor scavenger receptor BI: an old dog with new tricks. Curr Opin Endocrinol Diabetes Obes 2013; 20:124-131

 

Pharmacological Causes of Hyperprolactinemia

ABSTRACT

 

Hyperprolactinemia represents a multifaceted endocrine disorder with both physiological and pathological causes. The increased use of anti-psychotic and anti-depressant medications has increased the role pharmaceutical agents play in inducing hyperprolactinemia, being the most frequent cause of hyperprolactinemia in clinical practice. This has particularly impacted females, who demonstrate a higher susceptibility to drug-induced hyperprolactinemia. Of these medications, anti-psychotics, neuroleptic-like medications, anti-depressants, and histamine receptor type 2 antagonists, emerge as the most prominent culprits. Furthermore, opioids, some anti-hypertensive agents, proton pump inhibitors, estrogens, and other less potent hyperprolactinemia-inducing medications are recognized as potential contributors to drug-induced hyperprolactinemia. Many herbal medicines are reported as lactogenic, but their ability to cause hyperprolactinemia remains unclear. This review endeavors to elucidate the intricate mechanisms underlying the induction of hyperprolactinemia by pharmacological agents. We have included available data on the prevalence and extent of drug-induced changes in prolactin levels. We have also included data on herbal agents. We have highlighted where controversial data are identified. Although a detailed exploration of how these medications impact prolactin regulation is beyond the scope of this chapter, this review aims to deepen our understanding of the interplay between pharmacological agents and their effects on prolactin levels, contributing to valuable insights, refined therapeutic approaches, and better patient care.

 

INTRODUCTION

 

The most common cause of consistently high prolactin levels is drug-induced hyperprolactinemia. The overall incidence is higher in women compared to men. Drug-induced causes tripled during the 20-year follow-up period, reflecting the increased prevalence of psychoactive drug use (1). Several drugs have been reported to induce hyperprolactinemia, either by inhibiting dopamine receptors or their actions or by directly stimulating prolactin secretion (2).

 

High levels of prolactin can be attributed to various physiological factors, such as pregnancy and breast-feeding, while minor increases in prolactin levels may also occur during ovulation, after sexual intercourse, during periods of stress, exercise, after food intake, or in association with irritation of the chest wall and breast stimulation. Pathological causes can be related to hypothalamo-pituitary disorders or non-hypothalamo-pituitary disorders. Hypothalamo-pituitary disorders include prolactin-secreting pituitary tumors (including lactotroph tumors, mammo-somatotroph tumors, mature pluri-hormonal PIT1-lineage tumors, immature PIT1-lineage tumors, acidophil stem cell tumors, multi-hormonal pituitary tumors, mixed somatotroph, and lactotroph tumors, and pluri-hormonal tumors) (3); hypothalamic and pituitary stalk compression or damage (non-prolactin-secreting pituitary adenomas, craniopharyngiomas, meningiomas, germinomas, granulomas, metastasis, Rathke cleft cysts, hypophysitis, radiation, surgery, and trauma); infiltrative pituitary disorders; pituitary hyperplasia (McCune-Albright, Carney complex, X-LAG). Other causes include primary hypothyroidism; adrenal insufficiency; systemic diseases such as chronic renal failure and liver cirrhosis; polycystic ovary syndrome; neurogenic causes (chest wall trauma or surgery, herpes zoster); seizures; untreated severe phenylketonuria; pseudocyesis (false pregnancy); autoimmune diseases (lupus, rheumatoid arthritis, multiple sclerosis, systemic sclerosis, Behcet’s disease, polymyositis); cancers (breast, ovarian, colon, hepatocellular) (4). During the diagnostic process of hyperprolactinemia, it is crucial to consider the possibility of macroprolactinoma and the ‘hook’ effect, although the latter is usually not relevant in drug-induced cases (5).

 

Drug-induced hyperprolactinemia is often characterized by prolactin levels ranging from 25 to 100 ng/mL (530-2130 mIU/L). However, certain medications including metoclopramide, risperidone, amisulpride, and phenothiazines can lead to prolactin levels surpassing 200 ng/mL (4255 mIU/L) (6). On similar doses of prolactin-raising anti-psychotics, women with chronic use are more likely to develop hyperprolactinemia than men, reaching significantly higher prolactin levels, with mean levels of 50 ng/mL (1065 mIU/L) (7,8). Younger age was associated with higher prolactin levels in women, but not in men (8,9). Route of drug administration is important, with prolactin levels returning to normal after cessation of the drug: within 2-3 weeks after stopping oral treatment, but no sooner than 6 months after discontinuation of intramuscular depot administration (10).

 

This chapter will encompass a comprehensive discussion of all pharmacological causes as well as some alternative factors contributing to changes in prolactin levels especially hyperprolactinemia.

 

EPIDEMIOLOGY

 

Drug-induced hyperprolactinemia is the most common cause of consistently high prolactin levels. A retrospective follow-up study conducted in Scotland, involving 32,289 hyperprolactinemic individuals from 1993 to 2013, concluded that within the non-pregnancy-related group, the most prominent cause was drug-induced hyperprolactinemia (45.9%), followed by pituitary disorders (25.6%), macroprolactinoma (7.5%), and hypothyroidism (6.1%). Nevertheless, 15% of cases were deemed idiopathic. The overall incidence was higher in women aged 25-44 years old compared to men (1). Female predominance is reported in other studies with a female: male ratio of 5.9:1 and the mean age at diagnosis of hyperprolactinemia is 40 (range 14–85) years (2,11).The position of hyperprolactinemia as a side effect of medications has been assessed in a French Pharmacovigilance Database from 1985 to 2000, which reported 159 cases of hyperprolactinemia out of 182,836 adverse drug reactions (11). The rates of hyperprolactinemia related to therapeutic drug classes were recorded as 31% associated with anti-psychotics, 28% with neuroleptic-like drugs (medications with a similar mechanism of actions as neuroleptics, but used for different purposes, for example movement disorders or anti-emetics), 26% with anti-depressants, 5% with histamine receptor type 2 (H2-receptor) antagonists, and 10% with other drugs.

 

PROLACTIN CONTROL MECHANISMS

 

Prolactin is a polypeptide primarily produced in the anterior pituitary gland, with secondary production occurring in other tissues such as the gonads, mammary gland, endometrium, prostate, lymphocytes, hematopoietic cells, skin, brain, retina, inner ear cochlea, decidua, pancreas, liver, endothelium, and adipose tissue (12–14). In breast and prostate cancer, prolactin has even been proposed as a tumor marker (15,16). Prolactin acts through prolactin receptors (PRLR), which belong to the family of cytokine receptors associated with the non-receptor tyrosine Janus kinase 2. PRLR can activate the JAK-STAT (Janus kinase-signal transducer and activator of transcription) pathway, MAPK (mitogen-activated protein kinase), PI3 (phosphoinositide 3-kinase), Src kinase, as well as the Nek3 / Vav2 / Rac1 serine / threonine kinase pathway (17). There are different isoforms of this receptor: a long isoform, intermediate isoform, 2 short isoforms S1a and S1b which are formed by alternative splicing and partial deletion of exons 10 and 11, and soluble PRLR (18). These different isoforms are expressed in different tissues, mostly studied in rats. The long isoform is mainly expressed in the adrenal glands, kidneys, mammary glands, small intestine, bile ducts, choroid plexus, and pancreas whereas the short isoform is in the liver and ovaries (19). Prolactin possesses nearly 300 functions apart from lactation including neuroprotection and neurogenesis, offspring recognition by both parents, adipose and weight homeostasis, islet functions, immune regulation, angiogenesis, osmoregulation, and mitogenesis (20).

 

The secretion of prolactin produced by lactotroph cells in the anterior pituitary gland has a circadian rhythm with higher levels during sleep and lower levels during wakefulness (21). Even though pulsatility frequency does not significantly change over 24 hours, the amplitude of pulses is higher during night and day sleep, while wakefulness is associated with an immediate offset of active secretion. Prolactin is lower during the rapid eye movement stage of sleep (22).

 

The synthesis and secretion of prolactin is under the complex control of peptides, steroid hormones, and neurotransmitters, which can act as inhibitory or stimulatory factors, either by a direct effect on lactotroph cells or by indirect pathways through inhibition of dopaminergic tracts, and are widely studied in mammals (2). Dopamine plays a crucial role in inhibiting prolactin secretion. Dopamine can bind the five types of dopamine receptors (G-protein coupled receptors): DRD1, DRD2, DRD3, DRD4 and DRD5, while lactotroph cells express mainly D2 receptors. Dopamine can reach the pituitary via three pathways (Figure 1): through the tuberoinfundibular dopaminergic (TIDA) system, the tuberohypophysial tract (THDA), and the periventricular hypophyseal (PHDA) dopaminergic neurons (23). TIDA neurons originate from the rostral arcuate nucleus of the hypothalamus and release dopamine into the perivascular spaces of the medial eminence and through long portal vessels dopamine reaches the anterior pituitary gland. The THDA neurons originate in the rostral arcuate nucleus and project into the medial and posterior pituitary lobes and release dopamine at these sites. From THDA tract dopamine then reaches lactotroph cells through the short portal vessels (24). PHDA neurons originate in the periventricular nucleus and axons terminate in the intermediate lobe and dopamine release follows the same direction as from the THDA neurons. Prolactin-inhibiting neurons are considered to be a functional unit working synchronously (23). The binding of dopamine to D2 receptors on the plasma membrane of lactotroph cells inhibits prolactin protein, PRL gene transcription, as well as lactotroph proliferation (24). The release of prolactin through exocytosis of prolactin secretory granules is influenced by dopamine through various pathways. Specifically, D2 receptors are coupled with pertussis toxin-sensitive G proteins, which subsequently inhibit adenylate cyclase activity, resulting in decreased levels of cyclic adenosine monophosphate (cAMP) (25).

 

Additionally, the activation of potassium (K+) channels occurs, leading to a reduction in voltage-gated calcium (Ca2+) currents and inhibition of inositol phosphate production. Collectively, these intracellular signaling events culminate in a decrease in the concentration of free calcium ions (Ca2+) resulting in membrane hyperpolarization, ultimately inhibiting the exocytosis of prolactin from its granules (26). The inhibition of PRL gene transcription occurs when D2 receptors are activated, leading to the inhibition of MAPK or protein kinase C pathways. This activation results in a reduction of phosphorylation events on Ets family transcription factors. These transcription factors play a crucial role in the stimulatory responses of thyrotropin-releasing hormone (TRH), insulin, and epidermal growth factor (EGF) on prolactin expression. Moreover, the Ets family transcription factors interact with the PIT1 protein, which is essential for cAMP-mediated PRL gene expression (27). Dopamine exerts anti-mitogenic effects by activating D2 receptors through multiple pathways. These include the inhibition of MAPK (mitogen-activated protein kinase) signaling, protein kinase A signaling, and stimulation of phospholipase D activity. Additionally, dopamine engages a pertussis toxin-insensitive pathway, activates the extracellular signal-regulated kinases 1 and 2 (ERK1/2) pathway, and inhibits the AKT/protein kinase B pathway (28–31).

 

In addition to the dopaminergic inhibitory system, the γ-aminobutyric acid (GABA)-ergic tuberoinfundibular system, culminating in the median eminence, exhibits inhibitory properties, albeit of lesser potency compared to the dopaminergic system, while also having a role in prolactin modulation. GABA-B receptors are discernible both within the anterior pituitary gland, contributing to the maintenance of low prolactin levels, and in TIDA neurons which can be powerfully inhibited by GABA via hyperpolarization, consequently contributing to an elevation in prolactin levels (32,33).

 

Prolactin itself has two negative feedback effects: through ``short-loop feedback regulation`` it enhances the activity of TIDA neurons, where both long and short forms of the PRLR are expressed, with the long isoform being predominant in the arcuate and periventricular nuclei, regulating tyrosine hydroxylase (a rate-limiting enzyme in dopamine synthesis) leading to increase of dopamine release, which inhibits prolactin, as well as autocrine inhibition (34,35). Several other local factors influence prolactin release within pituitary gland as shown in Figure 1.

Figure 1. Prolactin – Central Nervous System Regulation.

 

Other prolactin inhibitory factors include somatostatin, acetylcholine, endothelins, gastrin, and growth hormone, while stimulatory factors include thyrotropin-releasing hormone (TRH) (as seen in primary hypothyroidism), angiotensin II, vasopressin, oxytocin, VIP, galanin, and estrogen.

 

Experiments conducted on rats to elucidate the relationship between the adrenergic system and the regulation of prolactin secretion have focused on stimulating or inhibiting α and β adrenergic receptors. Functional hyperprolactinemia is a complex hormonal interplay of stress-induced neuroendocrine changes involving the dopamine, serotonin and adrenergic systems (36,37). Evidence suggests that the mediobasal hypothalamus and preoptic-anterior hypothalamus harbor the primary adrenoreceptors (38,39). Injecting the α2 agonist clonidine into the mediobasal hypothalamus resulted in a dose-dependent increase in prolactin secretion. This effect was counteracted by the blockade of idazoxan (α2 antagonist). Similarly, the stimulation of prolactin release was induced by isoprenaline (β agonist) and notably attenuated by the β antagonist propranolol. The β2 agonist salbutamol also exhibited efficacy in stimulating prolactin secretion. Conversely, adrenergic agonists such as noradrenaline (mixed α and β), phenylephrine (α1), and tyramine (sympathomimetic) in the mediobasal hypothalamus, failed to elicit an effect on prolactin secretion.

 

Within the preoptic anterior hypothalamus, noradrenaline and adrenaline were found to stimulate prolactin secretion (40). However, the administration of the α1 agonist phenylephrine failed to stimulate prolactin, indirectly suggesting that the stimulatory effect of noradrenaline in the preoptic anterior hypothalamus is likely due to its action at α2 sites. α2 agonism has been shown to reduce the function of tuberoinfundibular dopaminergic neurons leading to increase prolactin production and secretion (41). Consequently, it was inferred that the activation of α2 and β adrenoceptors in the mediobasal hypothalamus and α2 adrenoceptors in the preoptic-anterior hypothalamus, proximal to prolactin-regulating neurons, leads to heightened prolactin secretion, while the action of α1 in the mediobasal hypothalamus may be inhibitory (42).

 

Cholinergic activation may have opposite roles in rodents and humans. Cholinergic agonists suppress prolactin release induced by morphine in rats, suggesting that the central cholinergic system has an inhibitory effect on the prolactin release triggered by morphine or β-endorphine, but this cholinergic inhibition does not occur through catecholaminergic neurons (43). Conversely, in humans, cholinomimetic drugs can increase prolactin levels associated with raised plasma β-endorphin, suggesting a stimulatory interplay of cholinergic factors and endogenous opioids on prolactin levels (44), although circulating opioids may not directly relate to central levels.

 

TIDA neurons express estradiol and progesterone receptors. Estradiol action leads to reduced secretion of dopamine into the portal blood system and mediates a prolactin surge. Progesterone, in addition, suppresses dopamine release being responsible for the plateau phase of the surge (23). Estrogen specifically affects prolactin synthesis by influencing lactotroph cell sensitivity, expression of pituitary dopamine receptor downregulation, and the expression of the prolactin receptor gene (2,34). Ghrelin, a hormone involved in metabolic balance, directly stimulates prolactin secretion at the pituitary level (45).

 

Tachykinins (substance P, neurokinins A, and B, neuropeptide K, neuropeptide ϒ) can act directly on the lactotroph cell and indirectly within the hypothalamus or posterior pituitary. They have a multifaceted impact on prolactin secretion, with both stimulatory and inhibitory effects. They can stimulate prolactin secretion by stimulating and potentiating the release of oxytocin, vasopressin, TRH, VIP, serotonin and glutamate, and by inhibiting GABA. Tachykinins through paracrine actions can directly increase prolactin within the anterior pituitary. They can also increase dopamine but the overall effect is prolactin elevation. Under specific circumstances, the stimulation of dopamine release can be prominent leading to a decrease in prolactin (46). Endogenous opioids are involved in regulating prolactin secretion, particularly during stressful situations, by reducing the activity of tuberoinfundibular dopaminergic neurons mediated by μ-, κ‑, and δ- opioid receptors, resulting in increased prolactin release (47). Prolonged nicotine exposure has been associated with desensitization of dopamine receptors, diminished dopamine turnover, and a decrease in their abundance within the nigrostriatal pathways (48). These alterations have been suggested to contribute to a diminished prolactin response to opiate blockade observed in individuals who smoke. Similar to opioids, histamine has been shown to induce prolactin production predominantly through inhibiting dopaminergic and stimulating serotoninergic and vasopressin-ergic neurons (49).

 

Serotoninergic pathways originating from the dorsal raphe nucleus play a physiological role in mediating nocturnal surges and suckling-induced prolactin rises through a serotonin interaction via serotonin type 1 and 2 receptors (5-hydroxytryptamine receptors, 5HT1, and 5HT2). 5-HT could either release a PRL-releasing factor or inhibit dopamine release. The paraventricular nucleus, where serotoninergic pathways terminate, contains postsynaptic serotonin 5-HT1A, 5-HT2, and 5-HT2C receptor subtypes, and possibly 5-HT3 receptors (50). It was shown that the prolactin-releasing effect of serotonin probably occurs mostly via 5-HT1C / 2 receptors because ritanserin (an elective 5-HT1C / 2 receptor antagonist) opposed this effect (51). Serotonin stimulation of prolactin–releasing factor (PRF) neurons in the paraventricular nucleus leads to PRF release (like VIP and oxytocin) mediating hyperprolactinemia. Moreover, serotoninergic stimulation of GABAergic neurons in the tuberoinfundibular-GABA system has been shown to inhibit TIDA cells which contain 5-HT1A receptors, therefore inhibiting dopamine synthesis/release resulting in increased prolactin secretion (52).

 

Oxytocin, through the posterior pituitary and vasoactive intestinal peptide (VIP) in the anterior pituitary, play significant roles in enhancing PRL gene transcription and modulating dopamine inhibition. Animal studies suggest a potential mediation of VIP by oxytocin to stimulate prolactin secretion (53).

 

The extensive hormonal regulation of prolactin renders it susceptible to various disturbances caused by different classes of medications.

 

CLINICAL CHARACTERISTICS

 

Persistent hyperprolactinemia is associated with disturbances of the gonadal axis leading to interruptions of gonadotrophin-releasing hormone pulsatility and inhibition of luteinizing hormone and follicle-stimulating hormone release (54). Clinical manifestations attributed to hyperprolactinemia predominantly stem from the suppression of the gonadal axis. In premenopausal women, a spectrum of menstrual cycle dysfunctions is observed, spanning from luteal phase shortening to complete amenorrhea, often correlating with elevated prolactin levels. Secondary amenorrhea can be due to hyperprolactinemia in up 30% of patients, and up to 75% of patients with amenorrhea and galactorrhea (55). Beyond these effects, an array of hypoestrogenic indicators may manifest, including symptoms like vaginal dryness, diminished libido, and decreased energy levels. Galactorrhea can be present in up to 80% of females (55,56). In men, the impact of hyperprolactinemia is manifested through a decrease in libido, ranging from diminished sexual desire to oligospermia or even azoospermia attributed to hypogonadotropic hypogonadism. Notably, erectile dysfunction may arise, primarily attributed to the direct inhibitory influence of dopamine, and can be potentially reversed through the administration of dopamine agonists (57). Gynecomastia, on the contrary, is a manifestation of secondary hypogonadism rather than elevated prolactin levels, whereas galactorrhea is rare in men (21).

 

In both genders infertility can be observed, with diminished bone mineral density. In females, bone mineral density is significantly decreased in women with amenorrhea and increases during treatment and menstrual cycle restoration (58). Additionally, in cases where hyperprolactinemia is attributed to a mass, accompanying clinical indications may encompass headaches, visual field disturbances, cranial nerve palsies, and hypopituitarism. Notably, these manifestations may be the only clinical features in post-menopausal women (21,59).

 

PSYCHOTROPIC MEDICATIONS

 

Anti-psychotics and neuroleptic-like drugs are psychotropic medications which primarily exert their anti-psychotic effects through the blockade of DRD2 and D4 receptors in the mesolimbic area. Newer classes of anti-psychotics block 5HT2 and sometimes noradrenergic α1 or α2 receptors (4,35,60). Blockade of D2 receptors in the hypothalamic tuberoinfundibular system and lactotroph cells results in disinhibition of prolactin secretion leading to hyperprolactinemia, being the most common drugs known to induce hyperprolactinemia (61) (Figure 2). On the contrary, strong binding to D2 receptors can extend the half-life of dopamine by approximately 50%. This effect is achieved through two primary mechanisms: direct blockade of the dopamine transporter (DAT) and antagonism of D2 autoreceptors. These processes collectively result in reduced reuptake of dopamine, prolonging its presence in the synaptic area and further stimulating an upregulation of receptors. However, it is important to note that chronic use of anti-psychotics can lead to the reversal of upregulation of DAT (mRNA and protein), potentially contributing to treatment resistance and potentially lower prolactin elevations in the long term. Nonetheless, it is worth mentioning that anti-psychotics typically exhibit a lower affinity for dopamine transporter blockade compared to selective DAT blockers such as nomifensine (62).

Figure 2. Mechanisms of drug-induced hyperprolactinemia with selected examples (adopted from La Torre et al. (2). In addition to opiates, cholinomimetics, PPIs and smoking indirectly also stimulate the opioid receptors. PPIs, Protein pump inhibitors; TCAs, tricyclic anti-depressants; MAO, monoamine oxidase; SSRIs, selective serotonin reuptake inhibitors; SNRIs, serotonin-noradrenaline reuptake inhibitors.

 

The potency of anti-psychotics and neuroleptic-like medications to induce a rise in prolactin levels varies (Table 1). The level of prolactin increase depends on the anti-psychotic drug (different affinity and selectivity for dopamine receptors; blood-brain barrier penetrating capability, degree of serotoninergic inhibition), the dose administered, and the patient's age and sex (34,35). Lastly, polymorphisms in genes related to dopamine receptors (such as DRD1, DRD2, DRD3) (63), dopamine transporters (SLC6A3), and dopamine-metabolizing enzymes (such as monoamine oxidase and catechol-O-methyltransferase) have been associated with individual variations in response to anti-psychotic treatment and the development of side effects, including hyperprolactinemia (64).

 

Although direct evidence establishing the involvement of adrenergic receptors in hyperprolactinemia caused by antipsychotic and antidepressant medications remains unproven, indirect indications, as elucidated in Figure 1, suggest the potential implication of these receptors. It is plausible that adrenergic receptors might play a partial role in the hyperprolactinemia induced by these medications.

           

Table 1. Medications and Their Ability to Cause Hyperprolactinemia

Cluster Name

Subclass mechanism of action

Medications

Prolactin increment

Frequency of prolactin increment (61,65)

Anti-psychotics

First generation anti-psychotics

Antagonize/block dopamine receptors, especially D2 receptors. Can block α1 adrenergic receptors.

Butaperazine

UP to 2-3-fold normal range with dozes 60 mg/daily. Higher in women (66)

High

Chlorpromazine

Up to 3-fold with initiation of treatment, up to 2-fold in long-term treatment) (67)

Moderate/ High

 

Flupenthixol

Up to 2-3-fold during the first month, and normalization in the next few months (68)

High

Fluphenazine

Up to 3-fold with initiation of treatment, up to 2-fold in long-term treatment (67). Up to 40-foldof the upper end of the normal range (69)

High

 

 

Haloperidol

Up to 9-fold at the beginning of treatment (3-fold in long-term treatment) (70)

High

Loxapine

Up to 3-foldof the upper end of the normal range in women (66)

Moderate

Perphenazine

Up to 40-fold of the upper end of the normal range (69)

Moderate

Pimozide

?

Moderate

Prochlorperazine

?

?

Promazine

Up to 4-fold of the upper end of the normal range (69)

 

Thiordiazine

Up to 3-fold with initiation of treatment, up to 2-fold in long-term treatment) (67)

High

Thiothixene

Up to 3-fold with initiation of treatment, up to 2-fold in long-term treatment (71)

Moderate/ High

 

Trifluoperazine

?

Moderate

Veralipride

Up to 10 time increment, transient (72)

High

Zuclopenthixole

?

?

Second generation anti-psychotics

Dopamine receptors blockade especially D2 receptors, serotonin (5-HT) receptor blockade, glutamate modulation, can antagonize α1 or α2 adrenergic receptors and histamine receptors.

Amisulpiride

 

Up to 10-fold at the beginning of treatment and remained elevated during treatment but lower levels (68)

Case reports

Aripiprazole

Reduce prolactin levels (73)

Case reports / No effect/ Reduced prolactin

Asenapine

Up to 2-fold increment and rarely with higher doses up to 4-fold (35)

Low, Moderate

 

 

Brexpiprazole

Mild increment (74)

 

Low

Clozapine

Mild (up to 2-fold) and transient (75)

 

Case reports or No effect

Iloperidone

Mild increment, transient (76)

Case reports or No effect

Levosulpiride

Up to 15-fold normal range (77)

Case reports / Moderate for galactorrhea (78)

 

Lurasidone

Up to 10-fold normal range (79)/ no effect (80)

Case reports or No effect

Molindone

?

Moderate

 

Olanzapine

Mild (up to 2-fold) and transient (75)

Low

 

Paliperidone

2-10-fold for depot formulations (81)

High

Perospirone

None (82)

None/ Case reports

Quetiapine

Mild and transient (75)

Low

 

Risperidone

2-10-fold

High (83)

 

Sertindole

Mild and transient (75)

?

Sulpiride

Up to 6-7-fold from baseline, dose dependent effect (84)

High

 

Thiethylperazine

?

?

Ziprasidone

Up to 4-fold from baseline and transient (35,75)

Low

 

Neuroleptic-like medications

 

Block D2 receptors

Domperidone

Up to 10-fold (85,86)

High

Droperidol

Significant increment after 10 minutes of administration, with peak at 20 minutes (87)

?

Metoclopramide

Up to 15-fold (2)

High

 

Anti-depressants

TCAs

Block the reuptake of both serotonin and noradrenaline.

Amitriptyline

2-fold increment on dosage 200/300mg (88)

Low

Amoxapine

3,5-fold to baseline (89)

High

Clomipramine

Up to 3-fold increment from baseline (90)

High

Desipramine

Just above the normal limit with 100 mg oral administration (91)

Low, Controversial

Imipramine

Up to 4-fold normal range (69)

Controversial

Nortriptyline

2-fold in the first 2 weeks in one patient (88)

None or Low

SSRI

Block the reuptake of serotonin.

Citalopram/

Escitalopram

Up to 3-fold increment (52)

None or Low (rare reports), Controversial data

Fluoxetine

Fluvoxamine

Paroxetine

Sertraline

SNRI

Block the reuptake of both serotonin and noradrenaline.

Duloxetine

Up to 2-fold normal range (92)

Case reports

Milnacipran

Not increased risk of hyperprolactinemia (93)

None

Venlafaxine

Up to 2-fold normal range, dose related (94)

Case reports

MAO inhibitors

Inhibit the enzyme. Monoamine oxidase, which breaks down serotonin, noradrenaline, and dopamine, though increasing their levels.

Clorgyline

Up to 2-fold from baseline (95)

Low

Pargyline

Up to 3-fold from baseline (95)

Low

Phenelzine

Unclear elevation, galactorrhea (96)

Low/ Case reports

Atypical anti-depressants

Inhibit noradrenaline and dopamine reuptake.

Bupropion

No significant change (80)

Case reports

Increases the release of both serotonin and noradrenaline.

Mirtazapine

No significant change (80)

Case reports

Serotonin modulators

Modulate serotonin receptors in the brain to enhance serotonin transmission.

Indoramine

 (97)

Case report

Nefazodone

Mild increment from baseline only at acute administration (98)

None/ Case reports

Trazodone

Up to1.5-fold from baseline (99)

None, Low

Vortioxetine

Up to 2-fold elevation (100)

Case reports

Selective noradrenaline reuptake inhibitor

Inhibit reuptake of norephinephrine.

Reboxetine

Up to 2-fold from baseline (101)

Case reports

NMDA receptor antagonist

Block NMDA receptors though influencing glutamate neurotransmission.

Esketamine

?

None

Gastric acid reducers

H2 receptor antagonists

H2 receptor antagonists.

Cimetidine

Up to 3-fold after 400 mg IV infusion (102)

Low

Ranitidine

Mild increment only in high IV doses (103)

Low

Protein pump inhibitors (PPIs)

Inhibit the activity of the proton pump (H+/K+ ATPase) in the stomach's parietal cells.

Esomeprazole

?

Case reports or No effect

Lansoprazole

4-fold increment from baseline (104)

Omeprazole

No significant change (105)

Pantoprazole

No significant change (106)

Rabeprazole

No significant change (107)

Opioids

 

They activate opioid receptors. Main types of opioid receptors: mu (μ), delta (δ), and kappa (κ).

Apomorphine

By acting as dopamine agonist it lowers prolactin (108)

None

Heroin

Elevated in addiction (within normal range) compared to healthy control or during abstinence (109)

Moderate in addicted patients that have values over 25 ng/mL

Methadone

Mild increment, transient increases for several hours following the administration (110)

?

Morphine

Up to 2-fold increment from baseline (111)

High

Antihypertensives

 

It decreases the release of noradrenaline.

Methyldopa

3-4-fold (65) up to 40-fold normal range (69)

Moderate

Inhibit the storage of neurotransmitters like noradrenaline and serotonin in nerve cells, though decreasing their release.

Reserpine

2.5-fold increment from baseline (112) Up to 40-fold normal range (69)

High

Block calcium channels in cardiac and smooth muscle cells.

Verapamil

2-fold (113)

Low

Estrogens

 

By using as contraceptives they suppress sexual axis.

Estradiol infusion

3-4-fold, dose-dependent, way of administration is important (oral and IV) (114)

Low

 

Estradiol withdrawal

?

 

Gonadotropins and GNRH agonists

 

Same as endogenous components, used for fertility induction.

hCG

Up to 4-fold increment, transient (115)

High

hMG

Up to 2.7-fold increment from baseline, transient (116)

High, Transient

GnRH agonist.

Leuprolide acetate

1.5-fold higher prolactin in compared to hMG alone, transient (117)

High

Other drugs

Benzodiazepines

Enhances the effects of GABA in the brain.

Diazepam

Mild, dose-dependent (118)

Controversial

Anxiolytics

Serotonin receptor agonist.

Buspirone

2-fold (119)

Case report or No effect

α-2 adrenergic agonist.

Clonidine

?

Case reports

Anticonvulsant

Block sodium channels in nerve cells.

Carbamazepine

Less than 2-fold in sleep entrained (120,121)

?

Phenytoin

Controversial, it can also lower prolactin levels (122)

?

Enhances the effects of GABA in the brain.

Phenobarbital

Controversial (123)

 

Valproic Acid

Controversial, it can also lower prolactin (124)

Case reports

Mood stabilizer

Decrease dopamine release and glutamate, increase GABA inhibition.

Lithium Carbonate

Controversial, no effect (183)

None

Antimigraine medication

Calcium channel blocker.

Flunarizine

Mild increment, up to 1.5-fold from baseline (125)

Case reports

Weight loss medications

Increase the release of serotonin and inhibit its reuptake.

Fenfluramine

Mild increment within normal range in previews non-hyperprolactinemic patients (126)

High

Inhibit the reuptake of serotonin, noradrenaline, and dopamine.

Sibutramine

4-fold (2,127)

Case report

Anticholinesterase inhibitors

Reversible acetylcholinesterase inhibitor.

Physostigmine salicylate

Less than 100 ng/mL (44).

Low

Prokinetic medication

Stimulate serotonin receptors in the gut.

Cisapride

High increment (up to 200 ng/dL) but in co-administration of other drug inducing hyperprolactinemia (128)

Case reports

Antihistaminic with sedative and antiemetic properties

Block histamine receptors.

Promethazine

?

?

Central Nervous System Stimulants

Increase the release and reduce the reuptake of noradrenaline and dopamine in the brain.

Amphetamine

Mild, only during withdrawal (129)

?

Methylphenidate

No effect (130)

Case reports/ No effect

ADHD medication

α -2 adrenergic agonist.

Guanfascine

Controversial, it can also lower prolactin (131)

Case reports

Decongestant

Sympathomimetic amine, predominantly α-1 agonist

Pseudoephedrine

Lower prolactin levels (132)

Case reports

Rheumatoid arthritis medications

Reduce inflammation, modify immune response.

Bucillamine

Mild increment within normal range (133)

Case report

Penicillamine

?

Case reports

Osteoporosis medication

Monoclonal antibody that inhibit the receptor activator of nuclear factor kappa-B ligand (RANKL).

Denosumab

?

Case reports

Substance of abuse

Blocks the reuptake of noradrenaline, dopamine, and serotonin in the brain.

Cocaine

Decrease prolactin levels (134)

Mild increment only during withdrawal (129)

Case reports

Increases the release and inhibits the reuptake of serotonin and to some extent, dopamine and noradrenaline.

Ecstasy

Mild or no effect (135)

?

Stimulates nicotinic acetylcholine receptors, leading to the release of neurotransmitters like dopamine and noradrenaline.

Smoking

Mild increment, transient (136)

Moderate

Anti-HIV medications

Protease inhibitors that prevent the cleavage of viral proteins and thereby inhibiting viral replication.

Ritonavir / Saquinavir

Mild (137)

Case reports

Radiotherapy

Use of high-energy radiation to damage the DNA within the targeted cells.

Intracranial radiotherapy

?

Moderate (138)

             

Frequency of increase to abnormal prolactin levels with chronic use: high: >50%; moderate: 25 to 50%; low: <25%; none or low: case reports. The effect may be dose-dependent. Drugs marked with blue have controversial data or decrease prolactin levels as explained in the table. Where we could not identify reliable data for the parameters in the table we added a question mark. *First-generation anti-psychotics, non-selective dopamine receptors antagonists. **Second-generation anti-psychotics.

 

Anti-Psychotics

 

Anti-psychotics are traditionally classified as first- and second-generation, but more recently a new classification taxonomy has been developed by McCutcheon et al. to express different receptor affinity of different anti-psychotics. Due to the impossibility to include in this new classification all drugs that cause hyperprolactinemia, we have used the old classification (Table 1) (139,140).

 

The first-generation anti-psychotics are typically associated with more severe hyperprolactinemia (2-3-fold increment), whereas second-generation drugs have lower D2 affinity and stronger blockade of 5HT2A receptors leading to milder prolactin elevations (1-2-fold), except risperidone, paliperidone, and amisulpiride. Amisulpiride has the greatest potential to cause hyperprolactinemia of all anti-psychotics (4).

 

The first-generation anti-psychotics, such as fluphenazine and haloperidol, act as non-selective dopamine receptors antagonists (2,10). The therapeutic effects on psychotic symptoms occur through D2 and D4 receptor binding in the mesolimbic area, while side effects are mediated by D2 blockade in the striatal area (linked to extrapyramidal effects) and in the hypothalamic infundibular system (linked to hyperprolactinemia) in more than 50% of patients. A clinical trial involving 69 patients examined the effects of various anti-psychotic medications on prolactin levels, including chlorpromazine, depot haloperidol, fluphenazine, zuclopenthixol, sulpiride, pimozide, droperidol, and flupenthixol. The study found a significant elevation in prolactin levels only in females, with a mean level of 1106 mIU/L (52 ng/mL) compared to the normal range of <480 mIU/L (22.6 ng/mL). In males, the mean prolactin levels were within the normal range, which may be attributed to the significantly lower total daily dose of chlorpromazine used in males (199.0-220.1 mg/day) compared to females (384.4-302.48 mg/day, P<0.05) (7).

 

Second-generation anti-psychotics with lower D2 affinity led to milder prolactin elevations (1-2-fold), except for paliperidone, risperidone, and amisulpiride whose effect on prolactin is similar to the first-generation neuroleptics. Chlorpromazine, loxapine, olanzapine and quetiapine have variable effects on prolactin secretion, while aripiprazole, clozapine, iloperidone, lurasidone have little or no effect on prolactin secretion (35).

 

An important factor contributing to variations in the induction of hyperprolactinemia by different anti-psychotic medications is the blood-brain barrier. Permeability glycoprotein transporter (P-gp), coded by the ABCB1 gene, is expressed in various tissues including in the cells of the blood-brain barrier. P-gp plays a role in actively transporting hydrophobic drugs with a molecular weight greater than 400 Da out of the brain, thus protecting the brain from these medications; therefore, this protein can change drug bioavailability (141,142).

 

The affinity of risperidone, paliperidone, and amisulpiride (prolactin rises up to 10-fold with these drugs) for P-gp is approximately twice that of olanzapine and chlorpromazine (prolactin rise is up to 3-fold with these drugs), and four times greater than haloperidol and clozapine (prolactin rise can be high initially but usually reduces with time) (143). The higher affinity of risperidone, paliperidone, and amisulpiride to P-gp could, among other mechanisms, partly explain the greater induction of hyperprolactinemia by these drugs, as P-gp does not allow them to enter the brain via the blood-brain barrier. Therefore, the portal circulation of the anterior pituitary delivers a somewhat higher concentration of these drugs to the lactotrophs, which are located outside the blood-brain barrier, to inhibit the D2 receptors (144).

 

Aripiprazole can act as a partial agonist at D2 receptors and display partial agonist activity at 5HT1A receptors, while also acting as an antagonist at 5HT2A receptors. Antagonism at these receptors can help to normalize prolactin levels since 5HT2A receptor activation has been associated with increased prolactin release. That is why it is considered a prolactin secretion modulator (145). Its role in prolactin levels has been investigated in a study involving both retrospective and prospective components (146). The retrospective part of the study included 30 patients undergoing risperidone treatment, when it was observed that after 6 months of treatment, prolactin levels remained high although somewhat lower than at the start of observation. In the prospective part of the study, 30 other patients were divided into two groups: one group receiving risperidone alone at a daily dosage of 2-4 mg and the other group receiving a combination of risperidone and aripiprazole at a daily dosage of 5-10 mg. The group receiving adjunctive aripiprazole exhibited significantly lower serum prolactin levels compared to the risperidone-only group at weeks1 (914±743 vs 1567±1009 mU/L), 2 (750±705 vs 1317±836 mUI/L) and 6 (658±590 vs 1557±882 mUI/L). Notably, during aripiprazole treatment, prolactin levels at weeks 1, 2, and 6 were significantly lower than at baseline (P< 0.05) (at baseline patients were treated with risperidone as monotherapy), suggesting that aripiprazole may effectively alleviate risperidone-induced hyperprolactinemia. Similar findings supporting the role of aripiprazole in reducing prolactin levels have been reported in other studies (147). Combination therapy presents a promising therapeutic approach for adjunctive treatment or for transitioning from risperidone to mitigate hyperprolactinemia (146).

 

More recently, a new medication SEP-363856, a trace amine-associated receptor 1 (TAAR1) and 5HT1A agonist, has been developed to treat schizophrenia. Its mechanism of action is not based on D2 antagonism, and has a favorable effectiveness and tolerability profile, without causing hyperprolactinemia (148). This category of medication serves as compelling evidence for the significant involvement of dopamine receptors in drug-induced hyperprolactinemia, and it is a future viable therapeutic choice for patients experiencing adverse effects associated with hyperprolactinemia.

 

DRUG-INDUCED HYPERPROLACTINEMIA IN PEDIATRIC PATIENTS

 

Anti-psychotic medications have been found to induce hyperprolactinemia in the pediatric population as well as in adults. In a trial involving 35 children and adolescents with early-onset psychosis, primarily diagnosed with childhood-onset schizophrenia or psychotic disorder not otherwise specified, prolactin levels were measured after a 3-week washout period, as well as after 6 weeks of treatment with haloperidol, olanzapine, and clozapine (149). Following the 6-week treatment period, haloperidol (9 of 10 patients – mean age 13.4 years) and olanzapine (7 of 10 patients– mean age 15.9 years) resulted in prolactin levels above the upper limit of normal. The mean increase was 5.2-fold for haloperidol and 2.4-fold for olanzapine. The prolactin response did not show statistically significant differences between females and males treated with haloperidol and olanzapine. Clozapine (22 patients, mean age 14.7) caused a small but significant rise in females (1.2-fold) but levels remained in the normal range for all patients. There was no rise in males. Why this difference between females and males is only on clozapine remains unclear and difficult to explain. However, in a study involving 36 girls aged 8-17 years, mean prolactin levels were higher in girls compared to boys, with the most significant increase occurring around the age of 13y, correlating with menarche. A highly significant correlation was found between increases in plasma prolactin and estradiol levels between the ages of 11 and 13 years. Girls with long menstrual cycles (>28 days) between the ages of 14 and 16 years had higher prolactin levels (p<0.05) (150). Even though the mean age of patients on clozapine was above 13 years, we do not possess information on the duration of the menstrual cycle of those girls, as if it is longer, the physiologic estrogenization because of the longer menstrual cycle can impact the range of prolactin elevation. In any case, the population sample size was relatively small to draw definitive conclusions and to provide answers why this happens only in clozapine patients (149).The authors concluded that the prolactin response in male children and adolescents treated with haloperidol or olanzapine was significantly higher than that observed in adult males. However, the prolactin response in female children and adolescents after haloperidol treatment did not differ significantly from that of adult females in similar studies, possibly due to the adult similarity of estrogen status seen in female adolescents (149). Aripiprazole in another study showed a lesser prolactin increase than olanzapine, quetiapine, and risperidone, similar to the adult population (151).

 

In another clinical trial involving 396 children and adolescents (aged 14.0 ± 3.1 years), the impact of anti-psychotic medications on prolactin levels was studied. The medications involved risperidone, olanzapine, quetiapine, and aripiprazole. Risperidone caused the highest incidence of hyperprolactinemia (93.5%) and had the highest peak prolactin levels (median = 56.1 ng/mL) followed in order by olanzapine, quetiapine, and aripiprazole. Menstrual disturbances were the most prevalent side effect (28.0%), particularly with risperidone (35.4%). Notably, severe hyperprolactinemia was associated with decreased libido, erectile dysfunction, and galactorrhea (152).

 

In conclusion, a comprehensive meta-analysis comprising 32 randomized controlled trials with a total of 4643 participants, with an average age of 13 years, has demonstrated that risperidone, paliperidone, and olanzapine are associated with a significant increase in prolactin levels among children and adolescents. Conversely, aripiprazole is linked to a notable decrease in prolactin levels in this age group. It is worth noting that haloperidol was not included in these studies, resulting in an absence of evidence regarding its prolactin-related effects in this population (153).

 

These findings underscore that haloperidol, risperidone, paliperidone and olanzapine are potent inducers of hyperprolactinemia in children and adolescents, mirroring observations in the adult population. A comprehensive listing of medications associated with hyperprolactinemia in children can be found in Table2.

 

Table 2. Drugs Reported to Induce Hyperprolactinemia in Children and Adolescents

Medication class

 

High

>50 percent of patients

Moderate

25-50 percent of patients

Low

<25 percent of patients

Case reports

Anti-psychotics, first-generation 'typical'

Fluphenazine (154)

Haloperidol (149,155)

 

Chlorpromazine (156)

Loxapine (157)

Pimozide (158,159)

 

 

Anti-psychotics, second-generation 'atypical'

Paliperidone (160,161)

Risperidone (152,155,162–164)

 

Asenapine (165)

Molindone (166)

Olanzapine (149,152)

Lurasidone (167,168)

Ziprasidone (169)

Quetiapine (152,162)

 

 

Clozapine (149)

Aripiprazole* (152,170)

Amisulpride (171)

Brexpiprazole (172)

Anti-depressants

Clomipramine (173)

 

 

Desipramine (174)

Bupropion (175)

Citalopram (176)

Escitalopram (177)

Fluoxetine (178)

Sertraline (179)

Duloxetine (177)

Paroxetine (180)

Venlafaxine (181)

Anti-emetics and gastrointestinal medications

Metoclopramide (182–184)

Domperidone (185,186)

 

 

 

Omeprazole (187)

Lansoprazole (187)

Cisapride

Others

Fenfluramine (188)

 

 

Estrogens (189)

Triptorelin (190)

Clonidine (191)

Methylphenidate (181)

Guanfacine (181)

Valproic acid (181)

Penicillamine (181)

*Aripiprazole is a partial agonist at the type 2 dopamine receptor and display partial agonist activity at the type 1A serotonin receptor (5HT1A) and antagonist at 5HT2A receptor. It can be used in combination with other psychotropic medications to reduce prolactin levels. Aripiprazole itself can sometimes cause mild hyperprolactinemia (192).

 

CLINICAL MANAGEMENT OF ANTI-PSYCHOTIC-INDUCED HYPERPROLACTINEMIA

 

Drug-induced hyperprolactinemia should be considered in the differential diagnosis of elevations of prolactin levels, sometimes greater than 200 µg/L (4260 mIU/L). Particularly when serum prolactin levels exceed 80-100 µg/L (1700-2130 mIU/L), pituitary magnetic resonance imaging (MRI) should be performed to rule out the presence of any underlying pituitary or hypothalamic masses that may contribute to hyperprolactinemia (193). According to the guidelines from the Endocrine Society, in symptomatic patients suspected of having drug-induced hyperprolactinemia, it is recommended the first test to diagnose drug-induced hyperprolactinemia is the discontinuation of the medication for 3 days or switch to an alternative drug (e.g. a prolactin-sparing anti-psychotic (e.g. aripiprazole), or an anti-psychotic with lower dopamine antagonist potency (Table 1) followed by retesting of serum prolactin levels.

 

However, any discontinuation or substitution of anti-psychotic agents should be done in consultation with the patient's psychiatric physician. If discontinuation is not possible or if the onset of hyperprolactinemia does not coincide with therapy initiation, obtaining a pituitary MRI is recommended (despite prolactin levels) to differentiate between medication-induced hyperprolactinemia and hyperprolactinemia caused by a pituitary or hypothalamic mass (194).

 

Before initiating treatment with an anti-psychotic medication, it is advised that clinicians inquire about the patient's previous treatment experience, sexual dysfunction, menstrual history (including irregularities and menopausal status), as well as any history of galactorrhea. Additionally, obtaining a baseline prolactin level is recommended before starting treatment (195). This pre-treatment screening for hyperprolactinemia can help determine whether subsequently elevated prolactin levels are due to medication-induced factors (196) and make the diagnosis easier without the need to perform further imaging.

 

While treatment of hyperprolactinemia in patients receiving anti-psychotics may not always be necessary, in cases where clinical hypogonadism is evident, several options are available (193). The initial step is to discontinue the drug if clinically feasible. If discontinuation is not possible, switching to a similar anti-psychotic that does not cause hyperprolactinemia is suggested. If neither of these options is feasible, cautious administration of a dopamine agonist may be considered in consultation with the patient's physician (194). It is worth noting that these interventions only result in the normalization of prolactin levels in approximately half of the patients receiving anti-psychotics, and careful psychiatric monitoring is required due to the possibility of psychosis exacerbation with dopamine agonists (193).

 

For patients with long-term hypogonadism demonstrated by hypogonadal symptoms or low bone mass, the use of estrogen or testosterone replacement is recommended. In rare instances where patients receiving anti-psychotics also present with a pituitary tumor, treatment options primarily revolve around tumor-specific interventions, considering especially optic chiasm compression. When a non-functioning tumor is suspected, the above options to eliminate the drug-induced component of hyperprolactinemia should be considered, with supervised short-term cessation of the medication to clarify drug-induced or tumor-derived hyperprolactinemia (193).

 

In addition to the interventions mentioned earlier, addressing fertility concerns in patients with drug-induced hyperprolactinemia may require further measures. Normalization of prolactin levels through medication adjustment or dopamine agonist therapy can often lead to the restoration of fertility. In situations where fertility is not spontaneously regained, fertility treatment with gonadotrophins or assisted reproductive procedures may be necessary.

 

NEUROLEPTIC-LIKE MEDICATIONS

 

Metoclopramide and domperidone are anti-emetic and gastrointestinal motility agents known also as neuroleptic-like medications which can increase pituitary prolactin secretion and breast milk production by a dopamine antagonistic action (Figure 1).

 

Metoclopramide is a central and peripheral D2 receptor antagonist (197). Its administration is followed by an acute increase in prolactin levels up to 15-fold above the baseline that persists in chronic administration of the drug (2). Even though the majority of patients treated with metoclopramide develop hyperprolactinemia (Table 1), related symptoms such as amenorrhea, galactorrhea, gynecomastia, and impotence remain unclear (2).

 

Domperidone is a peripheral D2 antagonist (it does not cross the blood-brain barrier) used for treating intestinal motility disorders, especially for the prevention of gastrointestinal discomfort with dopaminergic treatment in Parkinson's disease (198), as it antagonizes the D2 receptors in the upper gastrointestinal tract. It also reaches the D2 receptors on lactotroph cells inducing hyperprolactinemia, and can be used for stimulating lactation, for example in women with preterm infants, or an adoptive parent (199,200), and even in transgender women who wish to breastfeed (201,202). It can cause a 10-fold elevation in prolactin levels with normalization of prolactin levels after three days (85,86). Neuroleptic-like medications are summarized in Tables 1 and 3.

 

Table 3. H2 Receptor Antagonists, Opioids, Anti-Hypertensives, PPIs, Estrogens, and Other Drugs and Their Ability to Cause Hyperprolactinemia.

Medication class

High

>50% of patients

Moderate

25-50% of patients

Low

<25% of patients

Case reports

Anti-emetic and gastrointestinal

Domperidone

Metoclopramide

 

Prochlorperazine

 

Esomeprazole

Omeprazole

Lansoprazole

Cisapride

H2-receptor antagonists

 

 

Cimetidine

Ranitidine

 

Anti-hypertensives

 

Methyldopa

Verapamil

 

Others

Fenfluramine

Opioids

 

Estrogens

Protease inhibitors

Cocaine Bucillamine

Clonidine

Methylphenidate

Guanfascine

Valproic Acid

Penicillamine

 

ANTI-DEPRESSANTS

 

Anti-depressants can be classified based on their structure and mechanism of action into tricyclic anti-depressants (TCAs), selective serotonin reuptake inhibitors (SSRIs), serotonin-noradrenaline reuptake inhibitors (SNRIs), monoamine oxidase (MAO) inhibitors, atypical anti-depressants, serotonin modulators, selective noradrenaline reuptake inhibitor, and NMDA (N-Methyl-D-Aspartate) receptor antagonists (203). Data on their ability to cause hyperprolactinemia are controversial, as described below. The two main mechanisms, both related to elevated serotoninergic tonus, explain how anti-depressants can induce hyperprolactinemia by indirect modulation of prolactin release by serotonin and serotonin stimulation of GABAergic neurons (50) Adrenergic receptors involvement in their ability to cause hyperprolactinemia remains unclear. (Figure 1).

 

Data on the incidence of hyperprolactinemia with anti-depressant medications, especially SSRIs, MAO inhibitors, and some TCAs, suggest that they can cause modest and generally asymptomatic hyperprolactinemia (4,193). Their ability to cause hyperprolactinemia is summarized in Table 1.

 

TCAs

 

TCAs, such as amitriptyline, desipramine, clomipramine, and amoxapine, can induce sustained mild hyperprolactinemia (2) (Table 1). They manifest their mechanism of action by blocking the reuptake of noradrenaline and serotonin, through increasing serotoninergic tonus, leading to hyperprolactinemia. Amitriptyline's ability to cause hyperprolactinemia seems to be dose-dependent. A dosage of 150-250 mg/day to 5 patients showed no effect on prolactin levels after 3-7 weeks (204), whereas a dosage of 200-300 mg/day caused a 2-fold increment of prolactin levels in two of nine chronic treated patients (88). In 13 patients with depression taking amitriptyline or desipramine, prolactin levels were studied after intravenous injection of tryptophan (serotonin precursor). It was observed that tryptophan-induced prolactin elevation was significantly increased compared with a preceding placebo period (205). A similar tryptophan test was performed with clomipramine 20 mg vs placebo in 6 normal subjects. Levels of prolactin increased after tryptophan infusion in pretreated patients with clomipramine (206),suggesting serotonin involvement in this hyperprolactinemia.

 

The effect of the traditional TCA imipramine on serum prolactin levels is controversial. A five-week double-blind study on patients with depression taking imipramine did not show any significant change in serum prolactin (207). However, another study in young healthy men showed that imipramine’s effect on prolactin is dose-dependent (usually therapeutic dose is 50-150 mg): oral administration of 100 mg, but not of 40 mg, led to a consistent rise in prolactin levels after 3 hours of administration, but the rise was mild (maximum 25.85 ng/mL (550 mUI/L) (91).

 

Data regarding nortriptyline's ability to induce sustained hyperprolactinemia are lacking. Anyway, over a 4-6 week treatment of 8 patients with nortriptyline up to 150 mg/daily, no difference was found between placebo and the medication group. In only one patient was observed a transitional 2-fold elevation in the first 2 weeks (88).

 

Amoxapine, which is an anti-depressant with neuroleptic properties as well, is found to increase prolactin levels in 10 patients approximately 3.5-fold compared to baseline and more than desipramine in 12 patients, where almost no difference with baseline was observed (89). The proposed mechanism of hyperprolactinemia involves the blockade of the D2 receptor in tuberoinfundibular neurons or the anterior pituitary gland (2).

 

SSRIs

 

SSRIs enhance serotonin activity via inhibition of neuronal serotonin reuptake. This could be the most prominent mechanism leading to a prolactin elevation. A review of 13 case reports showed prolactin levels between 28 and 60 ng/mL (595 and 1276 mUI/L) (52). In a French study, 27 of 159 cases (17%) had SSRI-induced hyperprolactinemia with sertraline being the most prominent, followed by fluoxetine, paroxetine and fluvoxamine. Only citalopram was found not to increase prolactin levels significantly (208). However, in another study fluoxetine, paroxetine, and fluvoxamine were found again to elevate prolactin levels, but they also found citalopram to induce hyperprolactinemia. In this study duloxetine, milnacipran, and sertraline (which was the most prominent in the previous study) were not associated with an increased risk of hyperprolactinemia (93). Fluoxetine-induced hyperprolactinemia was found in patients with major depression (4.5% of men and 22.2% of women) following 12 weeks of fluoxetine treatment (209).Differences in the ability to cause hyperprolactinemia can be attributed to the variations in the affinity of the SSRIs for dopamine, histamine, and GABA receptors.

 

The Nurses’ Health Study and its follow-up study assessed anti-depressant use and circulating prolactin levels in 610 women (including 267 anti-depressant users) with two measurements of prolactin an average of 11 years apart (210). In this study, mean prolactin levels were similar among SSRI users (13.2 µg/L, (280 mUI/L), 95% CI 12.2-14.4), users of other classes of anti-depressants (12.7 µg/L (270 mUI/L), 95% CI 11.0-14.6), and non-users (13.1µg/L, (278 mUI/L), 95% CI 12.8-13.4) (210). However, the duration and dosage of anti-depressant use at the time of prolactin sampling had not been assessed, as the participants had only responded as current anti-depressant users/non-users on a questionnaire.

 

MAO Inhibitors

 

Monoamine oxidase is an enzyme responsible for breaking down neurotransmitters such as serotonin, noradrenaline, and dopamine in the brain. Inhibition of this enzyme is expected to increase the levels of all those neurotransmitters. Even though increased dopamine is suspected to be related to lower prolactin levels, probably the serotonin increment prevails and that is why this class of medications is related to hyperprolactinemia; or dopamine and serotonin increment neutralize each other and no difference in prolactin levels is seen.

 

MAO inhibitors with serotoninergic activity (pargyline and cordyline) can cause modest and generally asymptomatic hyperprolactinemia (2,61). Phenelzine was observed to persistently increase prolactin levels in 4 of11 patients, which returned to normal during a placebo week and rose again in all 4 patients after treatment restart (88).

 

Atypical Anti-Depressants

 

Mirtazapine's mechanism of action is different from other anti-depressants, as it does not inhibit the reuptake of serotonin or noradrenaline, but it increase the release of serotonin and noradrenaline (211). In any case, prolactin levels did not show any difference in 8 healthy male subjects pre- and post-mirtazapine 15 mg oral administration (212).

 

Serotonin Modulators

 

Trazodone acts through dual inhibition of serotonin reuptake and serotonin type 2 receptors, coupled with antagonism of histamine and α-1-adrenergic receptors (213). In 12 patients with depression, 150 mg of trazodone for 3 weeks caused significantly higher prolactin levels after 12 hours, 1 week, and 2 weeks of treatment compared to baseline. Higher levels were after the first week 15,3+/-8,5 ng/mL (325 +/- 180 mUI/L) compared to baseline 9,1 +/- 5,6 ng/mL (194+/-119 mUI/L) (99).

 

Selective Noradrenaline Reuptake Inhibitor

 

Reboxetine is a selective noradrenaline reuptake inhibitor that was shown to increase prolactin in healthy men after acute administration, but this effect can be reversed if reboxetine is simultaneously administered with α2-blocker mirtazapine, suggesting a role of α2-receptors in the enhancement of prolactin release after reboxetine (214). In any case, conflicting data are present even with this drug, with one other study showing no difference between pre-treatment and after-treatment prolactin levels in patients taking up to 8 mg reboxetine for 4 weeks in 17 patients (215). This discrepancy can be due to the limited number of patients or other neuroendocrine mechanisms still less explored.

 

SNRI medications are described to cause only mild and rare elevations on prolactin levels, whereas (94), esketamine (NMDA receptor antagonist)is not described to cause hyperprolactinemia.

 

Summary

 

In summary, controversial data are available on anti-depressant-induced hyperprolactinemia. Routine monitoring of prolactin levels in patients taking anti-depressants is not recommended unless symptoms related to prolactin increase (in premenopausal women: menstrual cycle dysfunction leading to amenorrhea, oligomenorrhoea, anovulatory cycles, low libido, and energy; in men: erectile dysfunction, decreased energy and libido, decreased muscle mass, decreased body hair; and for both of them osteopenia, galactorrhea and infertility) occur (50,193). If the prolactin serum level is elevated (> 25 µg/L (531 mUI/L), in these patients a differential diagnosis is needed. The proposed approach is to withdraw the anti-depressant drug slowly over 2 weeks and replace it with another anti-depressant less likely to cause hyperprolactinemia, then reassess symptoms and prolactin levels after 2-4 weeks (193). If the serum prolactin remains elevated, other causes of hyperprolactinemia should be addressed by an endocrinologist. Another approach is to perform a pituitary MRI if the replacement of the anti-depressant is difficult to manage.

 

GASTRIC ACID REDUCERS

 

Histamine-Receptor Inhibitors

 

Histamine, a CNS neurotransmitter, binds to both H1 and H2 receptors. It can stimulate prolactin secretion via H1 receptors by inhibiting the dopaminergic system. On the contrary, histamine can also inhibit prolactin secretion via H2 receptors using a non-dopaminergic mechanism involving β-endorphin, vasoactive intestinal peptide, vasopressin, or TRH (216), all of which act as prolactin-releasing factors (49) (Figure 1).

 

In the French Pharmacovigilance Study, H2 receptor antagonists were found to contribute to 5% of drug-induced hyperprolactinemia, primarily with ranitidine (odds ratio = 4.43; 95% CI: 1.82–10.8) (11). Other studies have shown that H2-receptor antagonists such as cimetidine and ranitidine can elevate prolactin levels (217,218). Specifically, cimetidine caused a three-fold increase in prolactin levels after a 400 mg IV infusion, although this effect was not observed with oral administration of 800 mg cimetidine in healthy individuals (102) (Table 1,3).

 

Interestingly, it has been observed that systemic administration of the H2 agonist impromidine does not prevent cimetidine-induced hyperprolactinemia. In contrast, pre-administration of benzodiazepines or GABA lowered the prolactin response. This suggests that cimetidine-induced hyperprolactinemia may be mediated through neurotransmitters in the GABA-ergic system (219).

 

Proton-Pump Inhibitors

 

Proton-pump inhibitors (PPIs) have been recently reviewed regarding the risk of hyperprolactinemia and related sexual disorders observed with long-term use (220). The exact mechanism by which PPIs increase prolactin levels is not fully understood; possible explanations include inhibition of dopamine receptors, interference with other dopamine receptors, involvement of the serotoninergic pathway, modulation of the opioid pathway, and a potential role in decreasing prolactin clearance (220) (Figure 2). In addition, PPIs can increase gastrin levels in chronic use especially in females using high doses (221). As gastrin can act as prolactin-inhibitory factor (108), this antagonistic effect may explain the relatively mild hyperprolactinemia occurring with this class of drugs. Esomeprazole has a mild inhibitory effect on CYP3A4, which leads to decreased metabolism of estrogen, thereby increasing serum estrogen levels which can stimulate the production of prolactin (222). Gynecomastia, impotence, irregular menses, and galactorrhea have been described with PPI use. Hyperprolactinemia occurred less often than sexual disorders, and most cases of hyperprolactinemia were reported with omeprazole, esomeprazole, and lansoprazole use (e.g. 4-fold increment with lansoprazole) (104) (Table 1,3). Pantoprazole and rabeprazole were only sporadically associated with hyperprolactinemia. The authors assert that the occurrence of sexual dysfunction in individuals using PPIs, despite having normal prolactin levels, may be attributed to the development of low vitamin B12 levels, hypomagnesaemia and iron deficiency resulting from PPI usage. These nutritional deficiencies have been implicated in the manifestation of sexual disorders in other studies (220,223,224).

 

OPIOIDS

 

Endogenous opioids, morphine, and related drugs, activate ε-, μ-, κ- and δ- opioid receptors in the hypothalamus, modulating pituitary hormone secretion. They do not possess direct effects on pituitary cells (225,226). They inhibit the gonadal axis via ε receptors (GnRH suppression) and stimulate prolactin production by reducing the activity of tuberoinfundibular dopaminergic neurons via μ, κ and δ opioid receptors – mostly μ receptors as the μ receptor opioid antagonist naloxone prior to morphine and methadone use prevents opioid-induced hyperprolactinemia (227). Indirect stimulation of prolactin release can be mediated by stimulating prolactin-releasing factors production (the serotoninergic pathways discussed above) (228) (Figure 2). Hyperprolactinemia can then lead to additional gonadal axis suppression. In addition, opioids can modulate the corticotroph axis via κ and δ opioid receptors and the somatotroph axis via μ, κ and δ opioid receptors (226). Additionally, opioids manifest negative effects on bone health through direct inhibition of osteoblasts by opioids, gonadal axis suppression, altered mental status, and other comorbidities (chronic conditions, smoking, alcohol use) (229).

 

Acute intravenous or intra-ventricular administration of endogenous opioids leads to a rapid plasma prolactin increase in a dose-dependent manner (44,225). Chronic use of opioids effects on prolactin can vary: oral opioids for chronic pain increase prolactin, but morphine administered intrathecally for chronic non-cancer pain had no effect on prolactin (226). Hyperprolactinemia induced by opioids can be symptomatic: painful gynecomastia, galactorrhea, and hypogonadism have been reported in chronic opioid users. These can be alleviated with discontinuation or reduction of opioid dose, and sometimes dopamine agonists such as bromocriptine (226). Methadone induces a transient increase in prolactin levels, whereas chronic methadone users have normal basal prolactin levels (110), (Table 1, 3).

 

On the contrary, in a study involving six patients with hyperprolactinemia and amenorrhea, the use of naltrexone, an opioid antagonist, was investigated to determine if blocking endogenous opioids could improve the sexual axis. On the first day of naltrexone administration, significant increases were observed in the mean concentration of luteinizing hormone (LH), LH pulse amplitude, and estradiol levels compared to the control day. This indicated a prompt partial reactivation of the hypothalamic-pituitary-gonadal axis as a result of naltrexone, leading to heightened gonadotrophin levels and subsequent release of estradiol. However, it was found that the effect of opioid antagonism did not result in a sustained increase in estradiol secretion with chronic treatment. Additionally, prolactin levels continued to increase over time (mean prolactin level 255 ± 121 microgram/L), despite the initial improvement in the gonadal axis. This study demonstrated that although prolactin-induced suppression of the gonadal axis can be reversed to some extent by acute opioid antagonism, it is not an effective treatment for revitalizing the gonadal axis in the long term. Possible explanations for this lack of sustained effect include desensitization of the hypothalamic-pituitary unit for the effects of opioid receptor blockade and other disruptors of the axis, which may counteract the positive effects of opioid antagonism (230,231).

 

For endocrinopathies caused by opioids, including hyperprolactinemia, potential management choices include reducing or discontinuing opioid usage whenever feasible and exploring alternative pain relief therapies for chronic pain situations. Hormonal replacement therapy can be considered for hypogonadism and hypoadrenalism (226).

 

ANTIHYPERTENSIVES

 

Some antihypertensive medications including α-methyldopa, verapamil, labetalol, and reserpine, have been associated with hyperprolactinemia. This phenomenon is attributed mainly to the potential inhibition of dopaminergic pathways, highlighting the complex interplay between anti-hypertensive therapy and endocrine function. Other mechanisms are drug-specific and will be explained below.

 

α -methyldopa is an α-adrenergic inhibitor that leads to the suppression of monoamine synthesis, including noradrenaline, dopamine, and serotonin, which likely contributes to its anti-hypertensive effect. It causes hyperprolactinemia through the inhibition of dopamine synthesis by competitive inhibition of DOPA decarboxylase which transforms L-dopa into dopamine (61) (Figure 2). Long-term treatment resulted in elevated basal prolactin levels (3-4-fold), while a single dose of 750-1000mg reaches a peak of high prolactin level after 4-6 hours of administration (232). Gynecomastia is the most common endocrine side effect.

 

Calcium channel blockers are other drugs studied for their potential to cause hyperprolactinemia. The dihydropyridine class was found to have no effects on prolactin levels. Whereas, from the non-dihydropyridine class, which mainly blocks L-type of calcium channel receptors in the heart, only verapamil was found to cause 2-fold persistent hyperprolactinemia (and galactorrhea),while drug discontinuation reversed hyperprolactinemia in all patients (113). A clinical trial suggested that verapamil acts by reducing dopamine release in the tuberoinfundibular pathway through calcium influx inhibition (Figure 2), possibly by N-calcium channels which are known to be involved in the regulation of dopamine release and other neurotransmitters (233).

 

Labetalol is an α- and β-adrenoceptor blocker anti-hypertensive that has been reported to increase prolactin levels when administered intravenously, but not when administered orally (100 or 200 mg) as labetalol cross the blood-brain-barrier only in negligible amounts. The increase in prolactin release caused by intravenous labetalol is not readily explained by its interference with adrenergic receptors. The exact mechanism underlying this effect of the drug is currently unclear, but it is possible that labetalol's ability to block dopamine activity (anti-dopaminergic activity) might be involved in this response (234) (Figure 2). Pre-treatment with levodopa and carbidopa can prevent prolactin response after labetalol (235), suggesting dopamine pathway involvement suppression inside the blood-brain-barrier.

 

Reserpine is a rauwolfia alkaloid previously used for the treatment of hypertension as well as psychosis, schizophrenia, and tardive dyskinesia; it reduces dopamine by inhibiting their hypothalamic storage in secretory granules (236), and by blockade of vesicular monoamine transporter type 2 in monoamine neurons (237), leading to hyperprolactinemia. Prolactin levels are higher during treatment with reserpine than 6 weeks after discontinuation of the drug. Increased incidence of gynecomastia and breast cancer has also been reported among patients on anti-hypertensive therapy with reserpine (236). The ability of anti-hypertensives to cause hyperprolactinemia is summarized in Tables 1 and 3.

 

ESTROGENS

 

Estrogens stimulate prolactin secretion by several mechanisms: They bind to specific intracellular lactotroph cells receptors, though enhancing prolactin gene transcription and synthesis (238). They also inhibit tuberoinfundibular dopamine synthesis, stimulate lactotroph cell hyperplasia, downregulate dopamine receptor expression, and modify lactotroph responsiveness to other regulators (23,239) (Figure 2). Estrogen-induced hyperprolactinemia is dependent on the degree of estrogenization. Higher levels of estrogens in pregnancy and during ovulation increase prolactin levels with the last, contributing to a higher normal range of prolactin in pre-menopausal women.

 

Studies documenting the incidence of hyperprolactinemia showed that women on oral contraceptives were reported to have higher prolactin levels by 12% to 30% (240,241). Some, but not all, studies suggest that there is a dose-dependent effect (4). No increase in basal prolactin levels is reported during therapy with modern contraceptives with lower amounts of estrogen (242)or estrogen plus cyproterone acetate alone (243).

 

In transgender patients, estradiol or ethinyl estradiol treatment, the prolactin level rise was dependent on the dose of estrogen, duration of exposure, and alteration of SHBG levels. Estradiol infusion at levels above 10,000 pg/mL for as short as 6-7 hours significantly elevated prolactin levels by 3- to 4-fold, whereas ethinyl estradiol 2 mg/day for 1 month did not consistently elevate prolactin in all patients, which can be due to its ability to increase SHBG binding and maintaining free portion in the normal range (114) (Table 1, 3).

 

For women on post-menopausal hormone replacement therapies over 2.5 years, serum prolactin measured were within the normal range (244). In another study on 75 women, who were randomly assigned to three groups: control (receiving placebo), transdermal hormonal replacement (biphasic 17β-estradiol and progesterone, natural hormones), and oral ethinyl-estradiol and desogestrel, prolactin levels significantly increased in the oral group, but not in the transdermal group. There was a significant difference in hormone levels: in the oral group, estradiol levels increased five times and estrone levels eleven times. In the transdermal group, estrone and estradiol levels were increased three times (245).

 

GONADOTROPHINS AND GNRH AGONISTS

 

In addition to the known prolactin function in lactation, several studies have suggested other benefits of prolactin in oocyte development, formation of corpus luteum and its survival, steroidogenesis and implantation (246). In natural cycles there is a transient increase in late follicular phase of prolactin, but this increment is higher in stimulated cycles (246). In a cohort study were included 79 patients; 60 individuals underwent in vitro fertilization, 14 received clomiphene citrate treatment, and five patients with premature ovarian failure were administered estradiol. During the course of human menopausal gonadotrophin (hMG) treatment, a notable increase in both serum estradiol and prolactin concentrations were observed from early to late follicular days (P < 0.01). Specifically, prolactin levels increased from an initial mean value of 367±38 mIU/L (17.25±1.8 ng/mL) to 991±84 mIU/L (46.6±4 ng/mL) (Table 1). Bromocriptine effectively mitigated the increase in prolactin levels but was associated with a significant elevation in estradiol levels (P < 0.05) because prolactin itself works as a controller of estradiol increment. Clomiphene treatment led to a significant increase in serum estradiol levels (P < 0.01) but a significant decrease in serum prolactin concentrations during the late follicular phase (P < 0.01), indicating disruption of the estradiol-prolactin feedback mechanism. Among patients with premature ovarian insufficiency, serum prolactin concentrations increased concomitantly with rising serum of estradiol concentrations (after estradiol administration). Additionally, it was observed that the presence of prolactin significantly reduced estradiol production by granulosa cells (P < 0.05) (116).

 

An increment of prolactin levels is found even after hCG administration with a maximum prolactin level of 93.2 ng/mL; 1983 mIU/L (115).Notably, knowing that prolactin is a stress hormone, during assisted procedures it is increased, but this is a transitory increment without consequences in fertility outcome (247).

 

Not only gonadotrophins but also GnRH agonists are widely used during invitro fertilization to maintain a controlled and synchronized ovarian stimulation. Use of leuprolide acetate (GnRH agonist) concomitantly with hMG, resulted in higher prolactin and estradiol levels in comparison with patients receiving only hMG (prolactin 24.2 vs 16.8 ng/mL; 515 vs 358 mIU/L) (117). In another randomized study, along protocol with 0.1 mg subcutaneous triptorelin starting from day 10 of the preceding stimulation cycle and short protocol, where 0.1 mg subcutaneous triptorelin is given in the stimulating cycle, were compared. Prolactin levels were measured at 9 am in the first day of hCG administration. The long protocol correlated with higher prolactin levels (31.3 ± 16.9 vs 23.7 ± 11 ng/mL; 666 ± 359 vs 504 ± 234 mIU/L) (248).

 

In children, GnRH agonists are used in precocious puberty (CPP)as well as growth hormone deficiency (GHD)who do not properly respond to exogenous growth hormone treatment. In a study involving 119 children with CPP and 93 with GHD, treated with triptorelin or leuprolide, prolactin levels were measured before and every six months for 6 years for CPP group and for 2 years for GHD group. Moreover, prolactin levels were checked after 6 and 12 months of treatment withdrawal. In this study was concluded that even though prolactin levels were higher in triptorelin treated patients (only 3.8% developed hyperprolactinemia in triptorelin group which was solved after withdrawal – baseline 12.5 ± 3.7 ng/mL (266 ± 79 mIU/L) to max 45.6 ± 4.5 ng/mL; 970 ± 96 mIL/L), no significant difference was found in prolactin in basal condition and during GNRH agonist treatment in CPP and GHD (190)(Table 2).

 

OTHER DRUGS

 

A lot of other drugs have been reported to cause mild (less than 2-fold increment) increases in prolactin levels. A synthesized visualization of these mechanisms is shown in Figure 2.

 

The acute administration of buspirone, an anxiolytic medication, was investigated in a study involving 8 healthy volunteers. The findings revealed an increase in plasma prolactin levels across all participants compared to the baseline levels observed in 8 control subjects. During the study, blood samples were collected at 30-minute intervals over a duration of 2 hours. The zenith of prolactin levels was observed between minutes 90 and 120 for all individuals, with the maximum elevation reaching 37 ng/mL (787 mIU/L) (249). It is noteworthy that the augmentation of prolactin is believed to exhibit a dose-dependent relationship. Furthermore, it was observed that chronic usage of buspirone did not lead to significant alterations in prolactin levels, indicating a potential adaptation to the acute changes induced by the medication. The underlying mechanism responsible for this phenomenon is posited to involve both serotoninergic and dopaminergic implications (119).

 

Carbamazepine, a widely used anticonvulsant, was examined in a cohort comprising 4 patients with complex partial seizures undergoing chronic carbamazepine treatment (200 mg administered three times daily). Blood samples were collected at intervals of 2 hours. Additionally, a group of 5 patients with untreated epileptic seizures participated, wherein a thyrotrophin-releasing hormone (TRH) stimulation test was performed both prior to and 35-50 days post the administration of 200 mg carbamazepine three times daily. Blood samples were obtained 10, 30, and 60 minutes following intravenous injection of 200µg TRH. Furthermore, 4 normal volunteer subjects were included in the study. On the first day, a placebo was administered, followed by the administration of 400 mg carbamazepine at 8 AM on the second day. Blood samples were collected at baseline on both days and subsequently at hourly intervals until 4 PM. After a span of two weeks, a nocturnal study was conducted, spanning from 6 PM to 6 AM. The investigation revealed that there were no discernible alterations in spontaneous prolactin release or TRH-stimulated prolactin levels. However, a slight increase in sleep-entrained prolactin values was observed, while retaining the secretory circadian rhythm. Given that the release of prolactin during sleep is largely attributed to serotoninergic activity, it is plausible that the modest increment (less than 2-fold) may implicate serotoninergic modulation (working as a serotonin-releasing factor and reuptake inhibitor) facilitated by carbamazepine (120,121).

 

Sympathomimetic amines fenfluramine and sibutramine, formerly used for appetite suppression due to their stimulatory effect on the synaptic concentration of serotonin, have been shown to induce hyperprolactinemia as a result of increased serotoninergic activity and postsynaptic stimulation of 5HT2Areceptors. In a case report, after starting sibutramine, a 38-year-old female patient developed hyperprolactinemia (prolactin levels 46 and 89.6 ng/mL (978 and 1906 mUI/L) with amenorrhea and galactorrhea. Discontinuation of sibutramine, confirmed by a sella MRI, led to rapid normalization of prolactin levels within 15 days, and symptoms resolved during a 90-day follow-up (2,127).

 

Cholinomimetic drugs have been reported controversially in the literature regarding their ability to cause hyperprolactinemia. However, in collaborative studies from the National Institute of Mental Health and the University of California, San Diego, three separate experiments were conducted involving volunteers of different genders and ages. In the first experiment, nine volunteers received physostigmine salicylate at 33 µg/kg, while in the second experiment, eleven male volunteers were given 22 µg/kg of physostigmine salicylate. The third experiment involved six volunteers receiving 3 mg of arecoline hydrobromide. Placebo saline was administered in all experiments as well. It was shown that intravenous injection of physostigmine or arecoline can elevate prolactin correlating with raised β-endorphin levels in the blood. Prolactin elevation was less than 100 ng/mL (2127 mUI/L). Cholinergic activation in the hypothalamus, particularly focusing on β-endorphin, might help in explaining how peptides modify primary neurochemical effects on hormone regulation in the hypothalamus and pituitary (44).

 

Bucillamine, an analogue of D-penicillamine used as an antirheumatic drug in Japan, has been reported to induce hyperprolactinemia (109 ng/mL (2319 mUI/L)) after 30 months of treatment start, associated with gynecomastia and galactorrhea in one case report. The mechanism remains unclear (133).

 

‘Ecstasy’ (MDMA) was shown to increase prolactin secretion in rhesus monkeys by stimulating serotonin release and by direct-acting as a 5HT2A agonist (250); In nine studies, five of them observed an increase in prolactin levels due to the intervention. However, in the remaining studies, there was no significant change in prolactin levels, and these unresponsive results tended to occur when a lower dose of the intervention was used on average. This suggests a potential relationship between the dosage of the intervention and its effect on prolactin levels (135).

 

Smoking, particularly the consumption of high-nicotine cigarettes, has been associated with a significant acute elevation in prolactin levels, ranging from 50% to 78% above the baseline, within 6 minutes after smoking. These elevated levels persist for approximately 42 minutes and return to baseline within 120 minutes of initiating smoking (136). The underlying mechanism probably involves the stimulation of rapid prolactin release through the augmentation of endogenous opioids, which subsequently inhibits dopamine release (251). However, prolonged nicotine exposure leads to desensitization of dopamine receptors, and lowers dopamine turnover (48) probably contributing to hyperprolactinemia. It has been hypothesized that the increased incidence of osteopenia and osteoporosis could be at least partly related to this effect (252).

 

Recently, an association has been reported between HIV-1 protease inhibitors and the adverse effect of galactorrhea and hyperprolactinemia in four HIV-1 infected women treated with indinavir, nelfinavir, ritonavir, or saquinavir. The cause of this unexpected toxicity could be attributed to several possible mechanisms: 1) Protease inhibitors may enhance the stimulatory effects of prolactin due to their inhibition of the cytochrome P450 system, leading to longer half-life of prolactin; 2) opportunistic infections in AIDS patients may induce cytokine-driven prolactin production by pituitary or immune cells; 3) protease inhibitors might exert direct endocrine effects on the pituitary or hypothalamus (253,254). To explore mechanisms of hyperprolactinemia induced by protease inhibitors, experiments were conducted using rat pituitary cells and hypothalamic neuronal endings. The results showed that both ritonavir and saquinavir could directly stimulate prolactin secretion, while not affecting dopamine release. This suggests that these protease inhibitors might interact with specific mammalian proteins in the anterior pituitary involved in prolactin secretion, leading to the observed galactorrhea and hyperprolactinemic effect (137).

 

Regarding chemotherapy and immunosuppression, there are some controversial data on the effect of chemotherapy and immunosuppression on prolactin levels, as significant prolactin increases are not frequent and usually mild (2). Prolactin and growth hormone have been involved as part of a cytokine system in the recovery of the immune response after chemotherapy and bone marrow transplantation (255). In a study of 20 breast cancer patients undergoing high-dose chemotherapy and autologous stem-cell transplantation, plasma prolactin levels increased within and 30 days after transplant, yet still remaining within the normal range. The use of antiemetic drugs further raised prolactin levels. Patients in continuous complete remission after transplantation exhibited higher prolactin levels, while elevated prolactin did not impact disease-free survival, suggesting potential for further research into post-transplant immune response (256).

 

Radiotherapy for intracranial germ cell tumors was shown to induce hyperprolactinemia with a prevalence of 35.3% (138).

 

Other drug inducing hyperprolactinemia are described in Table 1.

 

DRUGS REPORTED TO DECREASE PROLACTIN LEVELS OR HAVE AN EQUIVOCAL EFFECT

 

Several medications, beyond the established treatments like cabergoline, bromocriptine, carbidopa and levodopa, have reported effects on reducing prolactin levels. For instance, pseudoephedrine, an α-adrenergic stimulant primarily affecting α1 receptors, shares structural similarities with amphetamine and moderately stimulating dopamine release in the brain by acting on D2 receptors in the pituitary, consequently lowering prolactin levels. Studies have indicated pseudoephedrine's potential to decrease milk production, at least partly attributed to its effect on prolactin levels through dopaminergic actions in the pituitary (132). Moreover, indirect evidence suggests that α-1 receptors stimulation leads to decreased prolactin levels (41).

 

Amphetamine was seen to produce a poor prolactin suppressant effect in either normal- or hyperprolactinemic subjects. The proposed mechanism of prolactin lowering potential is due to their ability to stimulate the release of dopamine (257). However, during the withdrawal period of cocaine use, hyperprolactinemia has been observed, probably due to a decrease in dopamine levels, leading to dysregulation in the dopamine system and increased prolactin. Moreover, during withdrawal, prolactin can be secreted as a stress hormone (129).

 

Guanafascine, an α2 adrenergic agonist, used to treat ADHD, has been shown to decrease prolactin levels. In a longitudinal study spanning three years involving 15 patients diagnosed with hyperprolactinemia, the noteworthy suppressive impact of guanfacine on prolactin levels suggests potential involvement of hypothalamic or extrahypothalamic adrenergic pathways in the intricate regulation of prolactin secretion (131). Even though α2 stimulation has been shown to increase prolactin levels in rats, this is not fully understood in humans making the explanation in this case confusing (42).

 

The impact of benzodiazepines (BDZ) on prolactin secretion is a subject of debate. Research findings have yielded conflicting results. Some studies conducted on both non-epileptic patients and healthy volunteers have not detected significant modifications in prolactin levels following BDZ treatment (258). A study on 30 adolescent patients with schizophrenia with gradually increasing doses of diazepam to a maximum of 100-400 mg/day, with 4 weeks of treatment, showed that only doses higher than 250 mg/day give a significant but mild increase in prolactin levels. Proposed mechanisms are inhibition of TIDA neurons by activation of the GABA system, or activation of the endorphin-ergic system leading to hyperprolactinemia (118). On the contrary, diazepam was found to suppress the secretion of prolactin in vitro through one of two mechanisms: it either strengthens the direct inhibitory action of GABA on prolactin release, or it hinders a benzodiazepine-sensitive Ca2+-calmodulin dependent protein kinase at micromolar concentrations leading to a reduction of prolactin secretion (259).

 

Moreover, phenytoin, an anticonvulsant impeding sodium channels in nerve cells, have generated conflicting data regarding their impact on prolactin levels. In animal studies, phenytoin showcased a rapid decline in both prolactin release and mRNA concentrations, functioning as a partial T3 agonist by binding to T3 nuclear receptors (260). However, clinical observations revealed elevated resting levels of prolactin in phenytoin-treated patients compared to untreated counterparts. Remarkably, responses to metoclopramide and bromocriptine remained unaltered, indicating a limited effect of phenytoin on the D2 receptors present on lactotrophs (261). Even the conclusions drawn from these findings remain contentious. Evidence suggests that phenytoin treatment may enhance the growth hormone response to levodopa, implying a phenytoin-induced dopaminergic activity at the hypothalamic-pituitary level (122). More specifically, it is postulated that phenytoin might enhance dopamine receptor sensitivity by inhibiting the Ca2+ calmodulin complex. This effect could contribute to reduced prolactin secretion (122). On the contrary, other studies have not demonstrated any notable alterations in prolactin levels due to phenytoin administration (262). The discordant outcomes surrounding phenytoin's impact on prolactin levels underscore the complexity of its effects and necessitate further investigation for conclusive insights into its mechanisms of action.

 

In another comprehensive study involving 126 subjects, both males and females, with generalized or partial epilepsy receiving phenobarbital as monotherapy or in combination with phenytoin or benzodiazepines, a distinct pattern emerged. Specifically, the administration of phenobarbital, either alone or in combination, resulted in elevated prolactin levels, but this elevation was found to be statistically significant only in the male participants. Notably, knowing that an epileptic attack itself can cause hyperprolactinemia, those data remain confusing. The proposed mechanism in this study is phenobarbital interaction with GABA receptors, leading to increased prolactin levels (263). However other studies do not show any change in prolactin levels (123).

 

Valproic acid, an anticonvulsant working as a central stimulant of GABAergic neurons, has demonstrated the ability to reduce prolactin basal levels as well as TRH-stimulated prolactin levels. This is indirect proof of the synergically acting of GABA neurons with dopaminergic tracts (124). However, in another study, no effect of valproic acid on prolactin levels was noticed during the night (264).

 

Lithium carbonate, a pharmaceutical agent employed as a mood stabilizer, has undergone investigation in different studies, as it decreases dopamine release and glutamate, and increases inhibitory GABA (265). One of them encompassed a longitudinal examination involving 9 patients diagnosed with bipolar disorder. The focus of this study was the assessment of plasma prolactin levels before and 12 hours after the evening administration of lithium. Evaluations were conducted on days 1, 6, 8, 13, 30, 60, and 90. Notably, this investigation yielded no discernible correlation between lithium concentration and prolactin levels, and no statistically significant alterations in prolactin levels were observed. The second part of this study adopted a cross-sectional design, involving 26 patients with an established history of long-term lithium treatment spanning durations of 3 months to 20 years. A comparative analysis revealed that prolactin levels, measured at 9 AM following a one-hour period of rest, did not demonstrate elevation in comparison to 16 controls. It is noteworthy that in both studies, lithium concentrations ranged from 0.4 to 1.4 mmol/L (normal range 0.5-1.2 mmol/L) (266). Additionally, the administration of lithium did not exert an impact on the plasma prolactin response to thyrotrophin-releasing hormone (TRH) stimulation compared to pre-treatment levels (267). The combined findings from these investigations provide compelling evidence that lithium does not contribute to hyperprolactinemia, thereby distinguishing it from medications with such an effect.

Cocaine has been shown to decrease prolactin levels beginning at 30-min following cocaine administration reaching statistical significance at the 90- and 120-minute time points (134).

 

Those medications are mentioned in Table 1. Their mechanism of altering prolactin levels is summarized in Figure 2.

 

HERBAL MEDICINES AFFECTING PROLACTIN LEVELS

 

In Table 4 is list of herbal medicines has been used traditionally to stimulate lactation (268). However, firm scientific evidence that they actually induce hyperprolactinemia is scarce.

Table 4. Lactogenic Herbs (268)

Family name

Species name

Common name

Amaryllidaceae

Allium sativum

Garlic

Annonaceae

Xylopia aethiopica

 

African Pepper or Ethiopian Pepper

Asclepiadaceae

Secamoneafzelii

 

-

Costaceae

 

Costusafer

 

African Ginger

Euphorbiaceae

 

Euphorbia hirta

 

Asthma Plant or Tawa-Tawa

Euphorbia thymifolia

 

Petty Spurge

Hymenocardiaacida

 

African Almond or Honeytree

Plagiostylesafricana

 

-

Ricinus communis

 

Castor Bean Plant

Leguminosae

 

Tamarindus indica

 

Tamarind

Acacia nicolita

 

-

Desmodiumadscendens

 

-

Malvaceae

 

Hibiscus sabdariffa

 

Roselle or Red Sorrel

Gossypium herbaceum

 

Cotton Plant

Moraceae

 

 Milicia excelsa

 

African Teak or Iroko

Ficus species

Ficus or Fig trees

Musaceae

 

Musa paradisiaca

 

Plantain

Ranunculaceae

 

Nigella sativa

 

Black Cumin or Black Seed

 

Actaea (Cimiciguga) racemose

Black Cohosh

Solanaceae

 

Solanum torvum

 

Turkey Berry or Devil's Fig

Verbanaceae

 

Lippia multiflora

 

Bush Tea or False Green Tea

Zingiberaceae

 

Aframomummelegueta

 

Grains of Paradise or Alligator Pepper

Fabaceae

Trifolium pratense

Red Clover

Trigonella foenum-graecum

 

Fenugreek

Apiaceae

Foeniculum vulgare

Fennel

Some herbs are known to decrease prolactin levels. For example, chaste tree (Vitex agnus-castus) decrease prolactin levels by activating to D2-receptors and suppressing prolactin release, as shown in in vitro experiments on lactotroph cell cultures and in in vivo animal experiments (269). Another herb, Mucuna pruriens, which is a natural source of l-dihydroxyphenylalanine (a dopamine precursor) is found to decrease prolactin levels in humans (270). Vitamin B6 (pyridoxine), by acting as a coenzyme in dopamine synthesis and aspartame, a sweetener metabolized in phenylalanine (dopamine precursor), have been shown to interfere with milk production by reducing prolactin levels (271). Ashgawanda (Withania somnifera) is found to decrease prolactin levels up to 12% (272). Moreover, oral zinc is found to decrease prolactin levels below the normal range in all 17 subjects with normal prolactin levels, in scenario of increased zinc levels in the blood (273).

 

While none of the mentioned herbs are currently established within clinical guidelines for specifically lactogenic or prolactin-reducing purposes, ongoing research and anecdotal evidence suggest potential roles for these botanicals as adjunctive therapies.

 

CONCLUSION

 

In summary, this review underscores the significant role of drug-induced hyperprolactinemia in causing higher prolactin levels and provides detailed insights into how pharmaceutical agents contribute to this effect. However, understanding the complex mechanisms behind drug-induced hyperprolactinemia is still a work in progress. More research is needed to delve deeper into these mechanisms and gain better insights. These efforts will contribute to refining treatment strategies and improving patient care.

 

REFERENCES

 

  1. Soto-Pedre E, Newey PJ, Bevan JS, Greig N, Leese GP. The epidemiology of hyperprolactinaemia over 20 years in the Tayside region of Scotland: the Prolactin Epidemiology, Audit and Research Study (PROLEARS). Clin Endocrinol (Oxf). 2017 Jan;86 (1):60–7.
  2. Torre DL, Falorni A. Pharmacological causes of hyperprolactinemia. Ther Clin Risk Manag. 2007 Oct;3 (5):929–51.
  3. Asa SL, Mete O, Perry A, Osamura RY. Overview of the 2022 WHO Classification of Pituitary Tumors. Endocr Pathol. 2022 Mar 1;33 (1):6–26.
  4. Samperi I, Lithgow K, Karavitaki N. Hyperprolactinaemia. J Clin Med. 2019 Dec 13;8 (12):2203.
  5. Vilar L, Vilar CF, Lyra R, Freitas M da C. Pitfalls in the Diagnostic Evaluation of Hyperprolactinemia. Neuroendocrinology. 2019;109 (1):7–19.
  6. Ajmal A, Joffe H, Nachtigall LB. Psychotropic-induced hyperprolactinemia: a clinical review. Psychosomatics. 2014;55 (1):29–36.
  7. Smith S, Wheeler MJ, Murray R, O’Keane V. The effects of antipsychotic-induced hyperprolactinaemia on the hypothalamic-pituitary-gonadal axis. J Clin Psychopharmacol. 2002 Apr;22 (2):109–14.
  8. Kinon BJ, Gilmore JA, Liu H, Halbreich UM. Prevalence of hyperprolactinemia in schizophrenic patients treated with conventional antipsychotic medications or risperidone11This work was sponsored by Eli Lilly and Company. Psychoneuroendocrinology. 2003 Apr 1;28:55–68.
  9. Halbreich U, Kahn LS. Hyperprolactinemia and schizophrenia: mechanisms and clinical aspects. J Psychiatr Pract. 2003 Sep;9 (5):344–53.
  10. Wieck A, Haddad P. Hyperprolactinaemia caused by antipsychotic drugs. BMJ. 2002 Feb 2;324 (7332):250–2.
  11. Petit A, Piednoir D, Germain ML, Trenque T. Hyperprolactinémies d’origine médicamenteuse : étude cas/non-cas dans la banque nationale de pharmacovigilance. Therapies. 2003 Mar 1;58 (2):159–63.
  12. Kavarthapu R, Dufau ML. Prolactin receptor gene transcriptional control, regulatory modalities relevant to breast cancer resistance and invasiveness. Frontiers in Endocrinology [Internet]. 2022 [cited 2023 Jul 14];13. Available from: https://www.frontiersin.org/articles/10.3389/fendo.2022.949396
  13. Cabrera-Reyes EA, Limón-Morales O, Rivero-Segura NA, Camacho-Arroyo I, Cerbón M. Prolactin function and putative expression in the brain. Endocrine. 2017 Aug 1;57 (2):199–213.
  14. Marano RJ, Ben-Jonathan N. Minireview: Extrapituitary Prolactin: An Update on the Distribution, Regulation, and Functions. Mol Endocrinol. 2014 May;28 (5):622–33.
  15. Dagvadorj A, Collins S, Jomain JB, Abdulghani J, Karras J, Zellweger T, et al. Autocrine prolactin promotes prostate cancer cell growth via Janus kinase-2-signal transducer and activator of transcription-5a/b signaling pathway. Endocrinology. 2007 Jul;148 (7):3089–101.
  16. Jacobson EM, Hugo ER, Borcherding DC, Ben-Jonathan N. Prolactin in breast and prostate cancer: molecular and genetic perspectives. Discov Med. 2011 Apr;11 (59):315–24.
  17. Bouilly J, Sonigo C, Auffret J, Gibori G, Binart N. Prolactin signaling mechanisms in ovary. Molecular and Cellular Endocrinology. 2012 Jun 5;356 (1):80–7.
  18. Abramicheva PA, Smirnova OV. Prolactin Receptor Isoforms as the Basis of Tissue-Specific Action of Prolactin in the Norm and Pathology. Biochemistry Moscow. 2019 Apr 1;84 (4):329–45.
  19. Ben-Jonathan N, LaPensee CR, LaPensee EW. What can we learn from rodents about prolactin in humans? Endocr Rev. 2008 Feb;29 (1):1–41.
  20. Ben-Jonathan N, MERSHONf JL, Allen DL, Steinmetz RW. Extrapituitary Prolactin: Distribution, Regulation, Functions, and Clinical Aspects. 1996;17 (6).
  21. Prabhakar VKB, Davis JRE. Hyperprolactinaemia. Best Practice & Research Clinical Obstetrics & Gynaecology. 2008 Apr 1;22 (2):341–53.
  22. Spiegel K, Follenius M, Simon C, Saini J, Ehrhart J, Brandenberger G. Prolactin secretion and sleep. Sleep. 1994 Feb;17 (1):20–7.
  23. Grattan DR. 60 YEARS OF NEUROENDOCRINOLOGY: The hypothalamo-prolactin axis. J Endocrinol. 2015 Aug;226 (2):T101–22.
  24. Freeman ME, Kanyicska B, Lerant A, Nagy G. Prolactin: structure, function, and regulation of secretion. Physiol Rev. 2000 Oct;80 (4):1523–631.
  25. Gonzalez-Iglesias AE, Murano T, Li S, Tomić M, Stojilkovic SS. Dopamine Inhibits Basal Prolactin Release in Pituitary Lactotrophs through Pertussis Toxin-Sensitive and -Insensitive Signaling Pathways. Endocrinology. 2008 Apr 1;149 (4):1470–9.
  26. Vallar L, Meldolesi J. Mechanisms of signal transduction at the dopamine D2 receptor. Trends in Pharmacological Sciences. 1989 Feb 1;10 (2):74–7.
  27. Day RN, Liu J, Sundmark V, Kawecki M, Berry D, Elsholtz HP. Selective Inhibition of Prolactin Gene Transcription by the ETS-2 Repressor Factor *. Journal of Biological Chemistry. 1998 Nov 27;273 (48):31909–15.
  28. Booth AK, Gutierrez-Hartmann A. Signaling Pathways Regulating Pituitary Lactotrope Homeostasis and Tumorigenesis. In: Diakonova P Maria, editor. Recent Advances in Prolactin Research [Internet]. Cham: Springer International Publishing; 2015 [cited 2023 Jul 16]. p. 37–59. (Advances in Experimental Medicine and Biology). Available from: https://doi.org/10.1007/978-3-319-12114-7_2
  29. Suzuki S, Yamamoto I, Arita J. Mitogen-activated protein kinase-dependent stimulation of proliferation of rat lactotrophs in culture by 3’,5’-cyclic adenosine monophosphate. Endocrinology. 1999 Jun;140 (6):2850–8.
  30. Senogles SE. D2 dopamine receptor-mediated antiproliferation in a small cell lung cancer cell line, NCI-H69. Anticancer Drugs. 2007 Aug;18 (7):801–7.
  31. Radl D, Mei CD, Chen E, Lee H, Borrelli E. Each individual isoform of the dopamine D2 receptor protects from lactotroph hyperplasia. Molecular Endocrinology. 2013;27 (6):953–65.
  32. Grandison L, Cavagnini F, Schmid R, Invitti SC, Guidotti A. gamma-Aminobutyric acid- and benzodiazepine-binding sites in human anterior pituitary tissue. J Clin Endocrinol Metab. 1982 Mar;54 (3):597–601.
  33. Ammari R, Broberger C. Pre- and post-synaptic modulation by GABAB receptors of rat neuroendocrine dopamine neurones. Journal of Neuroendocrinology. 2020;32 (11):e12881.
  34. Fitzgerald P, Dinan TG. Prolactin and dopamine: what is the connection? A review article. J Psychopharmacol. 2008 Mar;22 (2 Suppl):12–9.
  35. Peuskens J, Pani L, Detraux J, De Hert M. The effects of novel and newly approved antipsychotics on serum prolactin levels: a comprehensive review. CNS Drugs. 2014 May;28 (5):421–53.
  36. de-Castro-e-Silva E, Ramalho MJ, Midlej M, Cobas C, Machado A, Antunes-Rodrigues J. Desipramine blocks stress-induced prolactin release in rats: role of central beta-2 adrenoceptors. Braz J Med Biol Res. 1990;23 (2):199–204.
  37. Sonino N, Navarrini C, Ruini C, Fallo F, Boscaro M, Fava GA. Life events in the pathogenesis of hyperprolactinemia. Eur J Endocrinol. 2004 Jul;151 (1):61–5.
  38. Willoughby JO, Day TA, Menadue MF, Jervois PM, Blessing WW. Adrenoceptors in the preoptic-anterior hypothalamic area stimulate secretion of prolactin but not growth hormone in the male rat. Brain Res Bull. 1986 May;16 (5):697–704.
  39. Day TA, Jervois PM, Menadue MF, Willoughby JO. Catecholamine mechanisms in medio-basal hypothalamus influence prolactin but not growth hormone secretion. Brain Res. 1982 Dec 16;253 (1–2):213–9.
  40. Koshiyama H, Kato Y, Shimatsu A, Murakami Y, Hattori N, Ishikawa Y, et al. Possible involvement of endogenous opioid peptides in prolactin secretion induced by alpha 2-adrenergic stimulation in rats. Proc Soc Exp Biol Med. 1989 Nov;192 (2):105–8.
  41. Fuxe K, Hökfelt T. Participation of Central Monoamine Neurons in the Regulation of Anterior Pituitary Function with Special Regard to the Neuro-Endocrine Role of Tubero-Infundibular Dopamine Neurons. In: Bargmann W, Scharrer B, editors. Aspects of Neuroendocrinology. Berlin, Heidelberg: Springer; 1970. p. 192–205.
  42. Kapoor R, Chapman IM, Willoughby JO. α2 and β Adrenoceptors in the Mediobasal Hypothalamus and α2 Adrenoceptors in the Preoptic-Anterior Hypothalamus Stimulate Prolactin Secretion in the Conscious Male Rat. Journal of Neuroendocrinology. 1993;5 (2):189–93.
  43. Muraki T, Tokunaga Y, Nakadate T, Kato R. Inhibition by cholinergic agonists of the prolactin release induced by morphine. Naunyn Schmiedebergs Arch Pharmacol. 1979 Sep;308 (3):249–54.
  44. Risch SC, Janowsky DS, Siever LJ, Judd LJ, Rausch JL, Huey LY, et al. Correlated cholinomimetic-stimulated beta-endorphin and prolactin release in humans. Peptides. 1982;3 (3):319–22.
  45. Lanfranco F, Motta G, Baldi M, Gasco V, Grottoli S, Benso A, et al. Ghrelin and anterior pituitary function. Front Horm Res. 2010;38:206–11.
  46. Debeljuk L, Lasaga M. Tachykinins and the control of prolactin secretion. Peptides. 2006 Nov;27 (11):3007–19.
  47. Shaw D, al’Absi M. Blunted opiate modulation of prolactin response in smoking men and women. Pharmacol Biochem Behav. 2010 Mar;95 (1):1–5.
  48. Netter P, Toll C, Lujic C, Reuter M, Hennig J. Addictive and nonaddictive smoking as related to responsivity to neurotransmitter systems. Behav Pharmacol. 2002 Sep;13 (5–6):441–9.
  49. Knigge U, Warberg J. Neuroendocrine functions of histamine. Agents Actions Suppl. 1991;33:29–53.
  50. Coker F, Taylor D. Antidepressant-induced hyperprolactinaemia. CNS Drugs. 2010 Jul 1;24 (7):563–74.
  51. Rittenhouse PA, Levy AD, Li Q, Bethea CL, Van de Kar LD. Neurons in the hypothalamic paraventricular nucleus mediate the serotonergic stimulation of prolactin secretion via 5-HT1c/2 receptors. Endocrinology. 1993 Aug;133 (2):661–7.
  52. Emiliano ABF, Fudge JL. From galactorrhea to osteopenia: rethinking serotonin-prolactin interactions. Neuropsychopharmacology. 2004 May;29 (5):833–46.
  53. Samson WK, Bianchi R, Mogg RJ, Rivier J, Vale W, Melin P. Oxytocin mediates the hypothalamic action of vasoactive intestinal peptide to stimulate prolactin secretion. Endocrinology. 1989 Feb;124 (2):812–9.
  54. Al-Chalabi M, Bass AN, Alsalman I. Physiology, Prolactin. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Aug 6]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK507829/
  55. Schlechte J, Sherman B, Halmi N, vanGilder J, Chapler F, Dolan K, et al. Prolactin-Secreting Pituitary Tumors in Amenorrheic Women: A Comprehensive Study*. Endocrine Reviews. 1980 Jul 1;1 (3):295–308.
  56. Vance ML, Thorner MO. Prolactinomas. Endocrinol Metab Clin North Am. 1987 Sep;16 (3):731–53.
  57. Carter JN, Tyson JE, Tous G, Van Vliet S, Faiman C, Friesen HG. Prolactin-Secreting Tumors and Hypogonadism in 22 Men. N Engl J Med. 1978 Oct 19;299 (16):847–52.
  58. Biller BM, Baum HB, Rosenthal DI, Saxe VC, Charpie PM, Klibanski A. Progressive trabecular osteopenia in women with hyperprolactinemic amenorrhea. The Journal of Clinical Endocrinology & Metabolism. 1992 Sep 1;75 (3):692–7.
  59. Thapa S, Bhusal K. Hyperprolactinemia. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Jun 19]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK537331/
  60. Montejo ÁL, Arango C, Bernardo M, Carrasco JL, Crespo-Facorro B, Cruz JJ, et al. Spanish consensus on the risks and detection of antipsychotic drug-related hyperprolactinaemia. Rev Psiquiatr Salud Ment. 2016;9 (3):158–73.
  61. Molitch ME. Medication-Induced Hyperprolactinemia. Mayo Clinic Proceedings. 2005 Aug 1;80 (8):1050–7.
  62. Amato D, Kruyer A, Samaha AN, Heinz A. Hypofunctional Dopamine Uptake and Antipsychotic Treatment-Resistant Schizophrenia. Front Psychiatry. 2019 May 28;10:314.
  63. Thompson J, Thomas N, Singleton A, Piggott M, Lloyd S, Perry EK, et al. D2 dopamine receptor gene (DRD2) Taq1 A polymorphism: reduced dopamine D2 receptor binding in the human striatum associated with the A1 allele. Pharmacogenetics. 1997 Dec;7 (6):479–84.
  64. Osmanova DZ, Freidin MB, Fedorenko OY, Pozhidaev IV, Boiko AS, Vyalova NM, et al. A pharmacogenetic study of patients with schizophrenia from West Siberia gets insight into dopaminergic mechanisms of antipsychotic-induced hyperprolactinemia. BMC Med Genet. 2019 Apr 9;20 (Suppl 1):47.
  65. Molitch ME. Drugs and prolactin. Pituitary. 2008 Jun 1;11 (2):209–18.
  66. Gruen PG, Sachar EJ, Altman N, Langer G, Tabrizi MA, Halpern FS. Relation of plasma prolactin to clinical response in schizophrenic patients. Arch Gen Psychiatry. 1978 Oct;35 (10):1222–7.
  67. Goodnick PJ, Santana O, Rodriguez L. Antipsychotics: impact on prolactin levels. Expert Opinion on Pharmacotherapy. 2002 Oct 1;3 (10):1381–91.
  68. Schlösser R, Gründer G, Anghelescu I, Hillert A, Ewald-Gründer S, Hiemke C, et al. Long-term effects of the substituted benzamide derivative amisulpride on baseline and stimulated prolactin levels. Neuropsychobiology. 2002;46 (1):33–40.
  69. Turkington RW. Prolactin Secretion in Patients Treated With Various Drugs: Phenothiazines, Tricyclic Antidepressants, Reserpine, and Methyldopa. Archives of Internal Medicine. 1972 Sep 1;130 (3):349–54.
  70. Spitzer M, Sajjad R, Benjamin F. Pattern of development of hyperprolactinemia after initiation of haloperidol therapy. Obstet Gynecol. 1998 May;91 (5 Pt 1):693–5.
  71. Crowley TJ, Hydinger-Macdonald M. Motility, Parkinsonism, and prolactin with thiothixene and thioridazine. Arch Gen Psychiatry. 1981 Jun;38 (6):668–75.
  72. Vercellini P, Sacerdote P, Trespidi L, Manfredi B, Panerai AE, Crosignani PG. Veralipride for hot flushes induced by a gonadotropin-releasing hormone agonist: a controlled study. Fertil Steril. 1994 Nov;62 (5):938–42.
  73. Cosi C, Carilla-Durand E, Assié MB, Ormiere AM, Maraval M, Leduc N, et al. Partial agonist properties of the antipsychotics SSR181507, aripiprazole and bifeprunox at dopamine D2 receptors: G protein activation and prolactin release. Eur J Pharmacol. 2006 Mar 27;535 (1–3):135–44.
  74. Ivkovic J, Lindsten A, George V, Eriksson H, Hobart M. Effect of Brexpiprazole on Prolactin: An Analysis of Short- and Long-Term Studies in Schizophrenia. J Clin Psychopharmacol. 2019;39 (1):13–9.
  75. Stanniland C, Taylor D. Tolerability of atypical antipsychotics. Drug Saf. 2000 Mar;22 (3):195–214.
  76. Weiden PJ. Iloperidone for the treatment of schizophrenia: an updated clinical review. Clin Schizophr Relat Psychoses. 2012 Apr;6 (1):34–44.
  77. Kuchay MS, Mithal A. Levosulpiride and Serum Prolactin Levels. Indian J Endocrinol Metab. 2017;21 (2):355–8.
  78. Lozano R, Concha MP, Montealegre A, de Leon L, Villalba JO, Esteban HL, et al. Effectiveness and safety of levosulpiride in the treatment of dysmotility-like functional dyspepsia. Ther Clin Risk Manag. 2007 Mar;3 (1):149–55.
  79. Suthar N, Aneja J. Lurasidone-induced Parkinsonism and Hyperprolactinemia. Indian J Psychol Med. 2019;41 (2):192–4.
  80. Pacchiarotti I, Murru A, Kotzalidis GD, Bonnin CM, Mazzarini L, Colom F, et al. Hyperprolactinemia and medications for bipolar disorder: Systematic review of a neglected issue in clinical practice. European Neuropsychopharmacology. 2015 Aug 1;25 (8):1045–59.
  81. Einarson TR, Hemels MEH, Nuamah I, Gopal S, Coppola D, Hough D. An analysis of potentially prolactin-related adverse events and abnormal prolactin values in randomized clinical trials with paliperidone palmitate. Ann Pharmacother. 2012 Oct;46 (10):1322–30.
  82. Togo T, Iseki E, Shoji M, Oyama I, Kase A, Uchikado H, et al. Prolactin levels in schizophrenic patients receiving perospirone in comparison to risperidone. J Pharmacol Sci. 2003 Mar;91 (3):259–62.
  83. Brunelleschi S, Zeppegno P, Risso F, Cattaneo CI, Torre E. Risperidone-associated hyperprolactinemia: evaluation in twenty psychiatric outpatients. Pharmacol Res. 2003 Oct;48 (4):405–9.
  84. Weizman A, Maoz B, Treves I, Asher I, Ben-David M. Sulpiride-induced hyperprolactinemia and impotence in male psychiatric outpatients. Prog Neuropsychopharmacol Biol Psychiatry. 1985;9 (2):193–8.
  85. da Silva OP, Knoppert DC, Angelini MM, Forret PA. Effect of domperidone on milk production in mothers of premature newborns: a randomized, double-blind, placebo-controlled trial. CMAJ. 2001 Jan 9;164 (1):17–21.
  86. Koch MW, Liu WQ, Camara-Lemarroy C, Zhang Y, Pike GB, Metz L, et al. Domperidone-induced elevation of serum prolactin levels and immune response in multiple sclerosis. J Neuroimmunol. 2019 Sep 15;334:576974.
  87. Schettini G, Mastronardi P, Scanni E, Pinto M, Forgione A, Florio T, et al. [Antiemetic effect of the levo isomer of sulpiride (L-sulpiride) in humans]. Minerva Anestesiol. 1989 May;55 (5):239–43.
  88. Meltzer HY, Fang VS, Tricou BJ, Robertson A. Effect of antidepressants on neuroendocrine axis in humans. Adv Biochem Psychopharmacol. 1982;32:303–16.
  89. Anton RF, Sexauer JD, Randall CL. Amoxapine elevates serum prolactin in depressed men. Journal of Affective Disorders. 1983 Nov 1;5 (4):305–10.
  90. Cordes J, Kahl KG, Werner C, Henning U, Regenbrecht G, Larisch R, et al. Clomipramine-induced serum prolactin as a marker for serotonin and dopamine turnover: results of an open label study. Eur Arch Psychiatry Clin Neurosci. 2011 Dec;261 (8):567–73.
  91. Nutt D, Middleton H, Franklin M. The neuroendocrine effects of oral imipramine. Psychoneuroendocrinology. 1987;12 (5):367–75.
  92. Özkan HM. [Duloxetine Associated Galactorrhea and Hyperprolactinemia: A Case Report]. Turk Psikiyatri Derg. 2020;31 (4):294–6.
  93. Trenque T, Herlem E, Auriche P, Dramé M. Serotonin Reuptake Inhibitors and Hyperprolactinaemia. Drug-Safety. 2011 Dec 1;34 (12):1161–6.
  94. Yang MS, Cheng WJ, Huang MC. Dose-related hyperprolactinemia induced by venlafaxine. Prog Neuropsychopharmacol Biol Psychiatry. 2009 Jun 15;33 (4):733–4.
  95. Slater SL, Lipper S, Shiling DJ, Murphy DL. Elevation of plasma-prolactin by monoamine-oxidase inhibitors. Lancet. 1977 Aug 6;2 (8032):275–6.
  96. Segal M, Heys RF. Inappropriate lactation. Br Med J. 1969 Oct 25;4 (5677):236.
  97. Pradalier A, Vincent D, Barzegar C. [Hyperprolactinemia induced by indoramin]. Therapie. 1998;53 (5):500–2.
  98. Ae W, Pj C. Attenuation of the prolactin-stimulating and hyperthermic effects of nefazodone after subacute treatment. J Clin Psychopharmacol. 1994;14:268–73.
  99. Otani K, Yasui N, Kaneko S, Ishida M, Ohkubo T, Osanai T, et al. Trazodone treatment increases plasma prolactin concentrations in depressed patients. Int Clin Psychopharmacol. 1995 Jun;10 (2):115–7.
  100. Ozkan HM. Galactorrhea and hyperprolactinemia during vortioxetine use: case report. Turkish Journal of Biochemistry. 2019 Feb 1;44 (1):105–7.
  101. Schüle C, Baghai T, Schmidbauer S, Bidlingmaier M, Strasburger CJ, Laakmann G. Reboxetine acutely stimulates cortisol, ACTH, growth hormone and prolactin secretion in healthy male subjects. Psychoneuroendocrinology. 2004 Feb;29 (2):185–200.
  102. Burland WL, Gleadle RI, Lee RM, Rowley-Jones D, Groom GV. Prolactin responses to cimetidine. Br J Clin Pharmacol. 1979 Jan;7 (1):19–21.
  103. Delitala G, Devilla L, Pende A, Canessa A. Effects of the H2 receptor antagonist ranitidine on anterior pituitary hormone secretion in man. Eur J Clin Pharmacol. 1982;22 (3):207–11.
  104. Duwicquet F, Gras-Champel V, Masmoudi K. Hyperprolactinémie avec galactorrhée induites par le lansoprazole : à propos d’un cas. Therapies. 2017 Dec 1;72 (6):691–3.
  105. MacGilchrist AJ, Howden CW, Kenyon CJ, Beastall GH, Reid JL. The effects of omeprazole on endocrine function in man. Eur J Clin Pharmacol. 1987;32 (4):423–5.
  106. Dammann HG, Bethke T, Burkhardt F, Wolf N, Khalil H, Luehmann R. Effects of pantoprazole on endocrine function in healthy male volunteers. Aliment Pharmacol Ther. 1994 Oct;8 (5):549–54.
  107. Dammann HG, Burkhardt F, Wolf N. The effects of oral rabeprazole on endocrine and gastric secretory function in healthy volunteers. Aliment Pharmacol Ther. 1999 Sep;13 (9):1195–203.
  108. Müller EE, Locatelli V, Cella S, Peñalva A, Novelli A, Cocchi D. Prolactin-Lowering and -Releasing Drugs Mechanisms of Action and Therapeutic Applications: Drugs. 1983 Apr;25 (4):399–432.
  109. Chan V, Wang C, Yeung RT. Effects of heroin addiction on thyrotrophin, thyroid hormones and porlactin secretion in men. Clin Endocrinol (Oxf). 1979 Jun;10 (6):557–65.
  110. Bart G, Borg L, Schluger JH, Green M, Ho A, Kreek MJ. Suppressed prolactin response to dynorphin A1-13 in methadone-maintained versus control subjects. J Pharmacol Exp Ther. 2003 Aug;306 (2):581–7.
  111. Delitala G, Grossman A, Besser GM. The participation of hypothalamic dopamine in morphine-induced prolactin release in man. Clin Endocrinol (Oxf). 1983 Oct;19 (4):437–44.
  112. Camanni E, Strumia E, Portaleone P, Molinatti GM. Prolactin secretion during reserpine and syrosingopine treatment. Eur J Clin Pharmacol. 1981;20 (5):347–9.
  113. Romeo JH, Dombrowski R, Kwak YS, Fuehrer S, Aron DC. Hyperprolactinaemia and verapamil: prevalence and potential association with hypogonadism in men. Clinical Endocrinology. 1996;45 (5):571–5.
  114. Goh HH, Ratnam SS. Effect of estrogens on prolactin secretion in transsexual subjects. Arch Sex Behav. 1990 Oct;19 (5):507–16.
  115. Kamel A, Halim AA, Shehata M, AlFarra S, El-Faissal Y, Ramadan W, et al. Changes in serum prolactin level during intracytoplasmic sperm injection, and effect on clinical pregnancy rate: a prospective observational study. BMC Pregnancy Childbirth. 2018 May 9;18 (1):141.
  116. Takiguchi S, Nakamura Y, Yamagata Y, Takayama H, Harada A, Sugino N, et al. Role of transient hyperprolactinemia in the late follicular phase of the gonadotropin‐stimulated cycle. Reprod Med Biol. 2002 Dec 11;1 (2):69–74.
  117. Meldrum DR, Cedars MI, Hamilton F, Huynh D, Wisot A, Marr B. Leuprolide acetate elevates prolactin during ovarian stimulation with gonadotropins. J Assist Reprod Genet. 1992 Jun 1;9 (3):251–3.
  118. Weizman A, Tyano S, Wijsenbeek H, Ben David M. High dose diazepam treatment and its effect on prolactin secretion in adolescent schizophrenic patients. Psychopharmacology (Berl). 1984;82 (4):382–5.
  119. Cowen PJ, Anderson IM, Grahame-Smith DG. Neuroendocrine effects of azapirones. J Clin Psychopharmacol. 1990 Jun;10 (3 Suppl):21S-25S.
  120. Bonuccelli U, Murialdo G, Martino E, Lecchini S, Bonura ML, Bambini G, et al. Effects of carbamazepine on prolactin secretion in normal subjects and in epileptic subjects. Clin Neuropharmacol. 1985;8 (2):165–74.
  121. Dailey JW, Reith ME, Steidley KR, Milbrandt JC, Jobe PC. Carbamazepine-induced release of serotonin from rat hippocampus in vitro. Epilepsia. 1998 Oct;39 (10):1054–63.
  122. D’Alessandro R, Cortelli P, Gallassi R, Ghisoli E, Montanaro N, Zoni E, et al. Phenytoin-induced increase in growth hormone response to levodopa in adult males. J Neurol Neurosurg Psychiatry. 1984 Jul;47 (7):715–9.
  123. Murialdo G, Galimberti CA, Gianelli MV, Rollero A, Polleri A, Copello F, et al. Effects of valproate, phenobarbital, and carbamazepine on sex steroid setup in women with epilepsy. Clin Neuropharmacol. 1998;21 (1):52–8.
  124. Giroud M, Marchal G, Baleydier-Giroud F, Desgres J, Dumas R. [GABAergic control of hypophyseal gonadotropic function and prolactin. Demonstration by valproic acid]. J Gynecol Obstet Biol Reprod (Paris). 1985;14 (2):163–9.
  125. Li W, Liu R, Liu W, Li G, Chen C. The effect of topiramate versus flunarizine on the non-headache symptoms of migraine. Int J Neurosci. 2023 Jan;133 (1):19–25.
  126. Scarduelli C, Mattei AM, Brambilla G, Zavaglia C, Adelasco P, Cavioni V, et al. Effect of Fenfluramine Oral Administration on Serum Prolactin Levels in Healthy and Hyperprolactinemic Women. Gynecol Obstet Invest. 1985 Feb 1;19 (2):92–6.
  127. Soares Leaes CG, Pereira-Lima JFS, da Costa Oliveira M. A case of sibutramine-induced hyperprolactinemia. Neuro Endocrinol Lett. 2011;32 (5):616–8.
  128. Ziambaras K, Dagogo-Jack S. Tumor-grade hyperprolactinemia induced by multiple medications in the setting of renal failure. Endocr Pract. 1999;5 (3):139–42.
  129. Lago JA, Kosten TR. Stimulant withdrawal. Addiction. 1994 Nov;89 (11):1477–81.
  130. Janowsky DS, Leichner P, Parker D, Judd L, Huey L, Clopton P. Methylphenidate and serum prolactin in man. Psychopharmacology. 1978 Jan 1;58 (1):43–7.
  131. Hauger-Klevene JH, Pinkas MB, Gerber S. Blood pressure and prolactin: effects of guanfacine. Three-year follow-up study. Hypertension. 1981;3 (6 Pt 2):II-222–5.
  132. Aljazaf K, Hale TW, Ilett KF, Hartmann PE, Mitoulas LR, Kristensen JH, et al. Pseudoephedrine: effects on milk production in women and estimation of infant exposure via breastmilk. Br J Clin Pharmacol. 2003 Jul;56 (1):18–24.
  133. Mori T, Yokogawa N, Higuchi R, Tsujino M, Shimada K, Sugii S. Bucillamine-induced gigantomastia with galactorrhea and hyperprolactinaemia. Mod Rheumatol Case Rep. 2020 Jan;4 (1):122–5.
  134. Elman I, Lukas SE. Effects of cortisol and cocaine on plasma prolactin and growth hormone levels in cocaine-dependent volunteers. Addictive Behaviors. 2005 May 1;30 (4):859–64.
  135. Dumont GJH, Verkes RJ. A review of acute effects of 3,4-methylenedioxymethamphetamine in healthy volunteers. J Psychopharmacol. 2006 Mar;20 (2):176–87.
  136. Mendelson JH, Sholar MB, Mutschler NH, Jaszyna-Gasior M, Goletiani NV, Siegel AJ, et al. Effects of Intravenous Cocaine and Cigarette Smoking on Luteinizing Hormone, Testosterone, and Prolactin in Men. J Pharmacol Exp Ther. 2003 Oct 1;307 (1):339–48.
  137. Orlando G, Brunetti L, Vacca M. Ritonavir and Saquinavir directly stimulate anterior pituitary prolactin secretion, in vitro. Int J Immunopathol Pharmacol. 2002;15 (1):65–8.
  138. Xiang B, Zhu X, He M, Wu W, Pang H, Zhang Z, et al. Pituitary Dysfunction in Patients with Intracranial Germ Cell Tumors Treated with Radiotherapy. Endocrine Practice. 2020 Dec 1;26 (12):1458–68.
  139. McCutcheon RA, Harrison PJ, Howes OD, McGuire PK, Taylor DM, Pillinger T. Data-Driven Taxonomy for Antipsychotic Medication: A New Classification System. Biol Psychiatry. 2023 Apr 14;S0006-3223 (23)01200-3.
  140. Chokhawala K, Stevens L. Antipsychotic Medications. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Jun 7]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK519503/
  141. Kelly RJ, Robey RW, Chen CC, Draper D, Luchenko V, Barnett D, et al. A pharmacodynamic study of the P-glycoprotein antagonist CBT-1® in combination with paclitaxel in solid tumors. Oncologist. 2012;17 (4):512.
  142. Schinkel AH. P-Glycoprotein, a gatekeeper in the blood-brain barrier. Adv Drug Deliv Rev. 1999 Apr 5;36 (2–3):179–94.
  143. Boulton DW, DeVane CL, Liston HL, Markowitz JS. In vitro P-glycoprotein affinity for atypical and conventional antipsychotics. Life Sci. 2002 May 31;71 (2):163–9.
  144. El-Mallakh RS, Watkins J. Prolactin Elevations and Permeability Glycoprotein. Prim Care Companion CNS Disord. 2019 May 16;21 (3):27149.
  145. Burris KD, Molski TF, Xu C, Ryan E, Tottori K, Kikuchi T, et al. Aripiprazole, a novel antipsychotic, is a high-affinity partial agonist at human dopamine D2 receptors. J Pharmacol Exp Ther. 2002 Jul;302 (1):381–9.
  146. Jiang XJ, Wu FX, Zhang JP, Shi L, Hu JQ, Zhu HZ, et al. Effects of Risperidone and Aripiprazole on Serum Levels of Prolactin, Testosterone and Estradiol in Female Patients with Schizophrenia. Drug Res (Stuttg). 2018 Jul;68 (7):410–4.
  147. Lee BJ, Lee SJ, Kim MK, Lee JG, Park SW, Kim GM, et al. Effect of Aripiprazole on Cognitive Function and Hyperprolactinemia in Patients with Schizophrenia Treated with Risperidone. Clin Psychopharmacol Neurosci. 2013 Aug 28;11 (2):60–6.
  148. Spoelstra SK, Bruggeman R, Knegtering H. [An antipsychotic without dopamine receptor blockade?]. Tijdschr Psychiatr. 2021;63 (11):804–9.
  149. Wudarsky M, Nicolson R, Hamburger SD, Spechler L, Gochman P, Bedwell J, et al. Elevated Prolactin in Pediatric Patients on Typical and Atypical Antipsychotics. Journal of Child and Adolescent Psychopharmacology. 1999 Jan;9 (4):239–45.
  150. Aubert ML, Sizonenko PC, Paunier L. Human Prolactin during puberty and first menstrual cycles in girls. Pediatr Res. 1978 Feb;12 (2):157–157.
  151. Uttley L, Kearns B, Ren S, Stevenson M. Aripiprazole for the Treatment and Prevention of Acute Manic and Mixed Episodes in Bipolar I Disorder in Children and Adolescents: A NICE Single Technology Appraisal. PharmacoEconomics. 2013 Nov 1;31 (11):981–90.
  152. Koch MT, Carlson HE, Kazimi MM, Correll CU. Antipsychotic-Related Prolactin Levels and Sexual Dysfunction in Mentally Ill Youth: A 3-Month Cohort Study. Journal of the American Academy of Child & Adolescent Psychiatry. 2023 Sep 1;62 (9):1021–50.
  153. Krøigaard SM, Clemmensen L, Tarp S, Pagsberg AK. A Meta-Analysis of Antipsychotic-Induced Hypo- and Hyperprolactinemia in Children and Adolescents. J Child Adolesc Psychopharmacol. 2022 Sep;32 (7):374–89.
  154. Bunker MT, Marken PA, Schneiderhan ME, Ruehter VL. Attenuation of antipsychotic-induced hyperprolactinemia with clozapine. J Child Adolesc Psychopharmacol. 1997;7 (1):65–9.
  155. Rosenbloom AL. Hyperprolactinemia with Antipsychotic Drugs in Children and Adolescents. Int J Pediatr Endocrinol. 2010;2010:159402.
  156. Apter A, Dickerman Z, Gonen N, Assa S, Prager-Lewin R, Kaufman H, et al. Effect of chlorpromazine on hypothalamic-pituitary-gonadal function in 10 adolescent schizophrenic boys. Am J Psychiatry. 1983 Dec 1;140 (12):1588–91.
  157. Hellings JA, Arnold LE, Han JC. Dopamine antagonists for treatment resistance in autism spectrum disorders: review and focus on BDNF stimulators loxapine and amitriptyline. Expert Opin Pharmacother. 2017 Apr;18 (6):581–8.
  158. Simeon J, Lawrence S, Simeon S. Effects of pimozide on prolactin in children [proceedings]. Psychopharmacol Bull. 1979 Apr 1;15 (2):40–2.
  159. Sallee FR, Dougherty D, Sethuraman G, Vrindavanam N. Prolactin monitoring of haloperidol and pimozide treatment in children with Tourette’s syndrome. Biol Psychiatry. 1996 Nov 15;40 (10):1044–50.
  160. Savitz A, Lane R, Nuamah I, Singh J, Hough D, Gopal S. Long-Term Safety of Paliperidone Extended Release in Adolescents with Schizophrenia: An Open-Label, Flexible Dose Study. J Child Adolesc Psychopharmacol. 2015 Sep;25 (7):548–57.
  161. Gopal S, Lane R, Nuamah I, Copenhaver M, Singh J, Hough D, et al. Evaluation of Potentially Prolactin-Related Adverse Events and Sexual Maturation in Adolescents with Schizophrenia Treated with Paliperidone Extended-Release (ER) for 2 Years: A Post Hoc Analysis of an Open-Label Multicenter Study. CNS Drugs. 2017 Sep 1;31 (9):797–808.
  162. Stevens JR, Kymissis PI, Baker AJL. Elevated prolactin levels in male youths treated with risperidone and quetiapine. J Child Adolesc Psychopharmacol. 2005 Dec;15 (6):893–900.
  163. Anderson GM, Scahill L, McCracken JT, McDougle CJ, Aman MG, Tierney E, et al. Effects of short- and long-term risperidone treatment on prolactin levels in children with autism. Biol Psychiatry. 2007 Feb 15;61 (4):545–50.
  164. Troost PW, Lahuis BE, Hermans MH, Buitelaar JK, van Engeland H, Scahill L, et al. Prolactin release in children treated with risperidone - Impact and role of CYP2D6 metabolism. Journal of Clinical Psychopharmacology. 2007 Feb;27 (1):52–7.
  165. Stepanova E, Grant B, Findling RL. Asenapine Treatment in Pediatric Patients with Bipolar I Disorder or Schizophrenia: A Review. Paediatr Drugs. 2018 Apr;20 (2):121–34.
  166. Stocks JD, Taneja BK, Baroldi P, Findling RL. A phase 2a randomized, parallel group, dose-ranging study of molindone in children with attention-deficit/hyperactivity disorder and persistent, serious conduct problems. J Child Adolesc Psychopharmacol. 2012 Apr;22 (2):102–11.
  167. Gjessing Jensen K. Severe Hyperprolactinemia during Lurasidone Treatment in a 16-year Old Girl with Schizophrenia - A Case Report. Scand J Child Adolesc Psychiatr Psychol. 2022 Jan;10 (1):87–8.
  168. FDA. Review and Evaluation of Clinical Data NDA #200603 [Internet]. 2019. Available from: https://www.fda.gov/media/133816/download
  169. Malone RP, Delaney MA, Hyman SB, Cater JR. Ziprasidone in adolescents with autism: an open-label pilot study. J Child Adolesc Psychopharmacol. 2007 Dec;17 (6):779–90.
  170. Yoo HK, Joung YS, Lee JS, Song DH, Lee YS, Kim JW, et al. A multicenter, randomized, double-blind, placebo-controlled study of aripiprazole in children and adolescents with Tourette’s disorder. J Clin Psychiatry. 2013 Aug;74 (8):e772-780.
  171. Varol Tas F, Guvenir T. Amisulpride treatment of adolescent patients with schizophrenia or schizo-affective disorders. Eur Child Adolesc Psychiatry. 2009 Aug;18 (8):511–3.
  172. FDA. Brexpiprazole in Pediatric Patients with Schizophrenia Aged 13 to 17 [Internet]. 2021. Available from: https://www.fda.gov/media/155985/download?attachment
  173. Hanna GL, McCracken JT, Cantwell DP. Prolactin in childhood obsessive-compulsive disorder: clinical correlates and response to clomipramine. J Am Acad Child Adolesc Psychiatry. 1991 Mar;30 (2):173–8.
  174. Overtoom CCE, Verbaten MN, Kemner C, Kenemans JL, van Engeland H, Buitelaar JK, et al. Effects of methylphenidate, desipramine, and L-dopa on attention and inhibition in children with Attention Deficit Hyperactivity Disorder. Behavioural Brain Research. 2003;145 (1–2):7–15.
  175. Conners CK, Casat CD, Gualtieri CT, Weller E. Bupropion hydrochloride in attention deficit disorder with hyperactivity. Journal of the American Academy of Child & Adolescent Psychiatry. 1996;35 (10):1314–21.
  176. Türkoğlu S. Citalopram-Induced Galactorrhea in an Adolescent. Clin Neuropharmacol. 2016;39 (6):331.
  177. McGrane IR, Morefield CM, Aytes KL. Probable Galactorrhea Associated with Sequential Trials of Escitalopram and Duloxetine in an Adolescent Female. J Child Adolesc Psychopharmacol. 2019 Dec;29 (10):788–9.
  178. Iancu I, Ratzoni G, Weitzman A, Apter A. More fluoxetine experience. J Am Acad Child Adolesc Psychiatry. 1992 Jul 1;31 (4):755–6.
  179. Ekinci N, Güneş S, Kalinli M, Ekinci Ö. Sertraline-Related Amenorrhea in an Adolescent. Clin Neuropharmacol. 2019;42 (3):99–100.
  180. Alacqua M, Trifirò G, Arcoraci V, Germanò E, Magazù A, Calarese T, et al. Use and tolerability of newer antipsychotics and antidepressants: a chart review in a paediatric setting. Pharm World Sci. 20070621st ed. 2008 Jan;30 (1):44–50.
  181. FDA. FDA Adverse Event Reporting System (FAERS) Public Dashboard. 2023.
  182. Ijaiya K. Prolactin response to exercise, metoclopramide and other provacative agents in children. Eur J Pediatr. 1980 Sep;134 (3):231–7.
  183. Preeyasombat C, Mahachoklertwattana P, Sriphrapradang A, Choubtam L. Prolactin (PRL) release in normal and growth hormone deficient children after oral metoclopramide. J Med Assoc Thai. 1993 Oct;76 Suppl 2:34–41.
  184. Nicolson R, Craven-Thuss B, Smith J, McKinlay BD, Castellanos FX. A Randomized, Double-Blind, Placebo-Controlled Trial of Metoclopramide for the Treatment of Tourette’s Disorder. Journal of the American Academy of Child and Adolescent Psychiatry. 2005 Jul 1;44 (7):640.
  185. Deprettere AR, Van Acker KJ, Du Caju MVL. Increased serum prolactin but normal TSH during prolonged domperidone treatment in children. Eur J Pediatr. 1987 Mar 1;146 (2):189–91.
  186. Cho E, Ho S, Gerber P, Davidson AGF. Monitoring of serum prolactin in pediatric patients with cystic fibrosis who are receiving domperidone. Can J Hosp Pharm. 2009 Mar;62 (2):119–26.
  187. Jabbar A, Khan R, Farrukh SN. Hyperprolactinaemia induced by proton pump inhibitor. J Pak Med Assoc. 2010 Aug;60 (8):689–90.
  188. Stoff DM, Pasatiempo AP, Yeung J, Cooper TB, Bridger WH, Rabinovich H. Neuroendocrine responses to challenge with dl-fenfluramine and aggression in disruptive behavior disorders of children and adolescents. Psychiatry Research. 1992 Sep 1;43 (3):263–76.
  189. Tack LJW, Heyse R, Craen M, Dhondt K, Bossche HV, Laridaen J, et al. Consecutive Cyproterone Acetate and Estradiol Treatment in Late-Pubertal Transgender Female Adolescents. J Sex Med. 2017 May;14 (5):747–57.
  190. Massart F, Parrino R, Placidi G, Massai G, Federico G, Saggese G. Prolactin secretion before, during, and after chronic gonadotropin-releasing hormone agonist treatments in children. Fertil Steril. 2005 Sep;84 (3):719–24.
  191. Mendhekar DN. Clonidine-induced gynecomastia and hyperprolactinemia in a 6-year-old child. J Clin Psychiatry. 2005 Dec;66 (12):1616–7.
  192. Koch MT, Carlson HE, Kazimi MM, Correll CU. Antipsychotic-Related Prolactin Levels and Sexual Dysfunction in Mentally Ill Youth: A 3-Month Cohort Study. Journal of the American Academy of Child & Adolescent Psychiatry. 2023 Sep 1;62 (9):1021–50.
  193. Chanson P. Treatments of psychiatric disorders, hyperprolactinemia and dopamine agonists. Best Practice & Research Clinical Endocrinology & Metabolism. 2022 Dec 1;36 (6):101711.
  194. Melmed S, Casanueva FF, Hoffman AR, Kleinberg DL, Montori VM, Schlechte JA, et al. Diagnosis and treatment of hyperprolactinemia: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab. 2011 Feb;96 (2):273–88.
  195. Peveler RC, Branford D, Citrome L, Fitzgerald P, Harvey PW, Holt RIG, et al. Antipsychotics and hyperprolactinaemia: clinical recommendations. J Psychopharmacol. 2008 Mar;22 (2 Suppl):98–103.
  196. Byerly M, Suppes T, Tran QV, Baker RA. Clinical implications of antipsychotic-induced hyperprolactinemia in patients with schizophrenia spectrum or bipolar spectrum disorders: recent developments and current perspectives. J Clin Psychopharmacol. 2007 Dec;27 (6):639–61.
  197. Isola S, Hussain A, Dua A, Singh K, Adams N. Metoclopramide. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Jun 20]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK519517/
  198. Lertxundi U, Domingo-Echaburu S, Soraluce A, García M, Ruiz-Osante B, Aguirre C. Domperidone in Parkinson’s disease: a perilous arrhythmogenic or the gold standard? Curr Drug Saf. 2013 Feb;8 (1):63–8.
  199. Asztalos EV, Campbell-Yeo M, da Silva OP, Ito S, Kiss A, Knoppert D, et al. Enhancing Human Milk Production With Domperidone in Mothers of Preterm Infants. J Hum Lact. 2017 Feb;33 (1):181–7.
  200. Glover K, Casey JJ, Gilbert M. Case Report: Induced Lactation in an Adoptive Parent. afp. 2023 Feb;107 (2):119–20.
  201. Weimer AK. Lactation Induction in a Transgender Woman: Macronutrient Analysis and Patient Perspectives. J Hum Lact. 2023 Aug;39 (3):488–94.
  202. Reisman T, Goldstein Z. Case Report: Induced Lactation in a Transgender Woman. Transgend Health. 2018;3 (1):24–6.
  203. Sheffler ZM, Patel P, Abdijadid S. Antidepressants. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Jun 9]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK538182/
  204. Meltzer HY, Piyakalmala S, Schyve P, Fang VS. Lack of effect of tricyclic antidepressants on serum prolactin levels. Psychopharmacology. 1977 Jan 1;51 (2):185–7.
  205. Charney DS, Heninger GR, Sternberg DE. Serotonin Function and Mechanism of Action of Antidepressant Treatment: Effects of Amitriptyline and Desipramine. Archives of General Psychiatry. 1984 Apr 1;41 (4):359–65.
  206. Anderson IM, Cowen PJ. Clomipramine enhances prolactin and growth hormone responses to L-tryptophan. Psychopharmacology (Berl). 1986;89 (1):131–3.
  207. Cooper DS, Gelenberg AJ, Wojcik JC, Saxe VC, Ridgway EC, Maloof F. The Effect of Amoxapine and Imipramine on Serum Prolactin Levels. AMA Archives of Internal Medicine. 1981 Jul;141 (8):1023–5.
  208. Petit A, Piednoir D, Germain ML, Trenque T. Hyperprolactinémies d’origine médicamenteuse : étude cas/non-cas dans la banque nationale de pharmacovigilance. Therapies. 2003 Mar 1;58 (2):159–63.
  209. Papakostas GI, Miller KK, Petersen T, Sklarsky KG, Hilliker SE, Fava M. Serum Prolactin Levels Among Outpatients With Major Depressive Disorder During the Acute Phase of Treatment With Fluoxetine. J Clin Psychiatry. 2006 Jun 15;67 (6):19359.
  210. Reeves KW, Okereke OI, Qian J, Tworoger SS, Rice MS, Hankinson SE. Antidepressant use and circulating prolactin levels. Cancer Causes Control. 2016 Jul;27 (7):853–61.
  211. Jilani TN, Gibbons JR, Faizy RM, Saadabadi A. Mirtazapine. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Aug 12]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK519059/
  212. Laakmann G, Schüle C, Baghai T, Waldvogel E. Effects of mirtazapine on growth hormone, prolactin, and cortisol secretion in healthy male subjects. Psychoneuroendocrinology. 1999 Oct;24 (7):769–84.
  213. Shin JJ, Saadabadi A. Trazodone. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 [cited 2023 Aug 12]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK470560/
  214. Schüle C, Baghai T, Laakmann G. Mirtazapine Decreases Stimulatory Effects of Reboxetine on Cortisol, Adrenocorticotropin and Prolactin Secretion in Healthy Male Subjects. Neuroendocrinology. 2004;79 (1):54–62.
  215. Moeller O, Hetzel G, Rothermundt M, Michael N, Nyhuis PW, Suslow T, et al. Oral citalopram and reboxetine challenge tests before and after selective antidepressant treatment. J Psychiatr Res. 2003;37 (3):261–2.
  216. Knigge UP. Histaminergic regulation of prolactin secretion. Dan Med Bull. 1990 Apr;37 (2):109–24.
  217. Perret G, Hugues JN, Louchahi M, Varoquaux O, Modigliani E. Effect of a short-term oral administration of cimetidine and ranitidine on the basal and thyrotropin-releasing hormone-stimulated serum concentrations of prolactin, thyrotropin and thyroid hormones in healthy volunteers. A double-blind cross-over study. Pharmacology. 1986;32 (2):101–8.
  218. Rampello L, Raffaele R, Nicoletti G, Pira FL, Vecchio I, Malaguarnera M, et al. Neurobehavioral Syndrome Induced by H2-Receptor Blocker Withdrawal: Possible Role of Prolactin. Clinical Neuropharmacology. 1997 Feb;20 (1):49–54.
  219. Sibilia V, Netti C, Guidobono F, Pagani F, Pecile A. Cimetidine-induced prolactin release: possible involvement of the GABA-ergic system. Neuroendocrinology. 1985 Mar;40 (3):189–92.
  220. Ashfaq M, Haroon MZ, Alkahraman YM. Proton pump inhibitors therapy and risk of hyperprolactinemia with associated sexual disorders. Endocrine Regulations. 2022 Apr 1;56 (2):134–47.
  221. Helgadóttir H, Lund SH, Gizurarson S, Metz DC, Björnsson ES. Predictors of gastrin elevation following proton pump inhibitor therapy. J Clin Gastroenterol. 2020 Mar;54 (3):227–34.
  222. Pipaliya N, Solanke D, Rathi C, Patel R, Ingle M, Sawant P. Esomeprazole induced galactorrhea: a novel side effect. Clin J Gastroenterol. 2016 Feb 1;9 (1):13–6.
  223. Bennett M. Vitamin B12 deficiency, infertility and recurrent fetal loss. J Reprod Med. 2001 Mar 1;46 (3):209–12.
  224. Toprak O, Sarı Y, Koç A, Sarı E, Kırık A. The impact of hypomagnesemia on erectile dysfunction in elderly, non-diabetic, stage 3 and 4 chronic kidney disease patients: a prospective cross-sectional study. Clin Interv Aging. 2017;12:437–44.
  225. Van Vugt DA, Meites J. Influence of endogenous opiates on anterior pituitary function. Fed Proc. 1980 Jun;39 (8):2533–8.
  226. Fountas A, Chai ST, Kourkouti C, Karavitaki N. Mechanisms in endocrinology: endocrinology of opioids. European Journal of Endocrinology. 2018 Oct 1;179 (4):R183–96.
  227. Panerai AE, Petraglia F, Sacerdote P, Genazzani AR. Mainly mu-opiate receptors are involved in luteinizing hormone and prolactin secretion. Endocrinology. 1985 Sep;117 (3):1096–9.
  228. Shin SH, Obonsawin MC, Van Vugt DA, Baby N, Jhamandas K. Morphine can stimulate prolactin release independent of a dopaminergic mechanism. Can J Physiol Pharmacol. 1988 Nov;66 (11):1381–5.
  229. Fountas A, Van Uum S, Karavitaki N. Opioid-induced endocrinopathies. Lancet Diabetes Endocrinol. 2020 Jan;8 (1):68–80.
  230. Matera C, Freda PU, Ferin M, Wardlaw SL. Effect of chronic opioid antagonism on the hypothalamic-pituitary-ovarian axis in hyperprolactinemic women. J Clin Endocrinol Metab. 1995 Feb;80 (2):540–5.
  231. de Wit W, Schoute E, Schoemaker J. Chronic naltrexone treatment induces desensitization of the luteinizing hormone pulse generator for opioid blockade in hyperprolactinemic patients. J Clin Endocrinol Metab. 1995 May;80 (5):1739–42.
  232. Steiner J, Cassar J, Mashiter K, Dawes I, Fraser TR, Breckenridge A. Effects of methyldopa on prolactin and growth hormone. Br Med J. 1976 May 15;1 (6019):1186–8.
  233. Kelley SR, Kamal TJ, Molitch ME. Mechanism of verapamil calcium channel blockade-induced hyperprolactinemia. Am J Physiol. 1996 Jan;270 (1 Pt 1):E96-100.
  234. Barbieri C, Ferrari C, Caldara R, Crossignani RM, Bertazzoni A. Endocrine and metabolic effects of labetalol in man. J Cardiovasc Pharmacol. 1981;3 (5):986–91.
  235. Barbieri C, Larovere MT, Mariotti G, Ferrari C, Caldara R. Prolactin Stimulation by Intravenous Labetalol Is Mediated Inside the Central Nervous System. Clinical Endocrinology. 1982;16 (6):615–9.
  236. Lee PA, Kelly MR, Wallin JD. Increased Prolactin Levels During Reserpine Treatment of Hypertensive Patients. JAMA. 1976 May 24;235 (21):2316–7.
  237. Veselinović T, Schorn H, Vernaleken IB, Schiffl K, Klomp M, Gründer G. Impact of different antidopaminergic mechanisms on the dopaminergic control of prolactin secretion. J Clin Psychopharmacol. 2011 Apr 1;31 (2):214–20.
  238. Frantz AG. Prolactin. N Engl J Med. 1978 Jan 26;298 (4):201–7.
  239. Arita J, Kimura F. Direct Inhibitory Effect of Long Term Estradiol Treatment on Dopamine Synthesis in Tuberoinfundibular Dopaminergic Neurons: In Vitro Studies Using Hypothalamic Slices. Endocrinology. 1987 Aug 1;121 (2):692–8.
  240. Reyniak JV, Wenof M, Aubert JM, Stangel JJ. Incidence of hyperprolactinemia during oral contraceptive therapy. Obstet Gynecol. 1980 Jan;55 (1):8–11.
  241. Luciano AA, Sherman BM, Chapler FK, Hauser KS, Wallace RB. Hyperprolactinemia and contraception: a prospective study. Obstet Gynecol. 1985 Apr;65 (4):506–10.
  242. Jernström H, Knutsson M, Taskila P, Olsson H. Plasma prolactin in relation to menstrual cycle phase, oral contraceptive use, arousal time and smoking habits. Contraception. 1992 Dec;46 (6):543–8.
  243. Acién P, Mauri M, Gutierrez M. Clinical and hormonal effects of the combination gonadotrophin-releasing hormone agonist plus oral contraceptive pills containing ethinyl-oestradiol (EE) and cyproterone acetate (CPA) versus the EE-CPA pill alone on polycystic ovarian disease-related hyperandrogenisms. Hum Reprod. 1997 Mar;12 (3):423–9.
  244. Foth D, Römer T. Prolactin serum levels in postmenopausal women receiving long-term hormone replacement therapy. Gynecol Obstet Invest. 1997;44 (2):124–6.
  245. Stanosz S, Zochowska E, Safranow K, Sieja K, Stanosz M. Influence of modified transdermal hormone replacement therapy on the concentrations of hormones, growth factors, and bone mineral density in women with osteopenia. Metabolism. 2009 Jan;58 (1):1–7.
  246. Iancu ME, Albu AI, Albu DN. Prolactin Relationship with Fertility and In Vitro Fertilization Outcomes—A Review of the Literature. Pharmaceuticals. 2023 Jan;16 (1):122.
  247. Trikoilis Ν, Mavromatidis G, Tzafetas M, Deligeoroglou Ε, Tzafetta M, Loufopoulos A, et al. The association of in vitro fertilization/intracytoplasmic sperm injection results with anxiety levels and stress biomarkers: An observational, case-control study✰. Journal of Gynecology Obstetrics and Human Reproduction. 2022 Jan 1;51 (1):102254.
  248. Kamel MA, Zabel G, Bernart W, Neulen J, Breckwoldt M. Comparison between prolactin, gonadotrophins and steroid hormones in serum and follicular fluid after stimulation with gonadotrophin-releasing hormone agonists and human menopausal gonadotrophin for an in-vitro fertilization programme. Hum Reprod. 1994 Oct;9 (10):1803–6.
  249. Meltzer HY, Flemming R, Robertson A. The Effect of Buspirone on Prolactin and Growth Hormone Secretion in Man. Archives of general psychiatry. 1983 Oct;40 (10):1099–102.
  250. Murnane KS, Kimmel HL, Rice KC, Howell LL. The neuropharmacology of prolactin secretion elicited by 3,4-methylenedioxymethamphetamine (“ecstasy”): A concurrent microdialysis and plasma analysis study. Horm Behav. 2012 Feb;61 (2):181–90.
  251. Xue Y, Domino EF. Tobacco/Nicotine and Endogenous Brain Opioids. Prog Neuropsychopharmacol Biol Psychiatry. 2008 Jul 1;32 (5):1131–8.
  252. Li H, Wallin M, Barregard L, Sallsten G, Lundh T, Ohlsson C, et al. Smoking-Induced Risk of Osteoporosis Is Partly Mediated by Cadmium From Tobacco Smoke: The MrOS Sweden Study. J Bone Miner Res. 2020 Aug;35 (8):1424–9.
  253. Hutchinson J, Murphy M, Harries R, Skinner CJ. Galactorrhoea and hyperprolactinaemia associated with protease-inhibitors. Lancet. 2000 Sep 16;356 (9234):1003–4.
  254. Montero A, Fernandez MA, Cohen JE, Luraghi MR, Sen L. Prolactin levels in the cerebrospinal fluid of patients with HIV infection and AIDS. Neurol Res. 1998 Jan;20 (1):2–4.
  255. Redelman D, Welniak LA, Taub D, Murphy WJ. Neuroendocrine hormones such as growth hormone and prolactin are integral members of the immunological cytokine network. Cell Immunol. 2008;252 (1–2):111–21.
  256. Hinterberger-Fischer M, Ogris E, Kier P, Bauer K, Kittl E, Habertheuer KH, et al. Elevation of plasma prolactin in patients undergoing autologous blood stem-cell transplantation for breast cancer: is its modulation a step toward posttransplant immunotherapy? Am J Clin Oncol. 2000 Aug;23 (4):325–9.
  257. DeLeo V, Cella SG, Camanni F, Genazzani AR, Müller EE. Prolactin lowering effect of amphetamine in normoprolactinemic subjects and in physiological and pathological hyperprolactinemia. Horm Metab Res. 1983 Sep;15 (9):439–43.
  258. Wilson JD, King DJ, Sheridan B. Tranquillisers and plasma prolactin. Br Med J. 1979 Jan 13;1 (6156):123–4.
  259. Schettini G, Cronin MJ, O’Dell SB, MacLeod RM. The benzodiazepine agonist diazepam inhibits basal and secretagogue-stimulated prolactin release in vitro. Brain Research. 1984 Jan 23;291 (2):343–9.
  260. Davis JR, Lynam TC, Franklyn JA, Docherty K, Sheppard MC. Tri-iodothyronine and phenytoin reduce prolactin messenger RNA levels in cultured rat pituitary cells. J Endocrinol. 1986 Jun;109 (3):359–64.
  261. Elwes RD, Dellaportas C, Reynolds EH, Robinson W, Butt WR, London DR. Prolactin and growth hormone dynamics in epileptic patients receiving phenytoin. Clin Endocrinol (Oxf). 1985 Sep;23 (3):263–70.
  262. Dana-Haeri J, Oxley J, Richens A. Pituitary responsiveness to gonadotrophin-releasing and thyrotrophin-releasing hormones in epileptic patients receiving carbamazepine or phenytoin. Clin Endocrinol (Oxf). 1984 Feb;20 (2):163–8.
  263. Bonuccelli U, Murialdo G, Rossi G, Bonura ML, Polleri A, Murri L. Prolactin Secretion in Epileptic Subjects Treated with Phenobarbital: Sex Differences and Circadian Periodicity. Epilepsia. 1986;27 (2):142–8.
  264. Cavallo A, Moore DC, Nahori A, Beaumanoir A, Sizonenko PC. Plasma prolactin and cortisol concentrations in epileptic patients during the night. Arch Neurol. 1984 Nov;41 (11):1179–82.
  265. Malhi GS, Tanious M, Das P, Coulston CM, Berk M. Potential mechanisms of action of lithium in bipolar disorder. Current understanding. CNS Drugs. 2013 Feb;27 (2):135–53.
  266. Lanng Nielsen J, Amdisen A, Darling S, Pedersen EB. Plasma prolactin during lithium treatment. Neuropsychobiology. 1977;3 (1):30–4.
  267. Tanimoto K, Maeda K, Yamaguchi N, Chihara K, Fujita T. Effect of lithium on prolactin responses to thyrotropin releasing hormone in patients with manic state. Psychopharmacology (Berl). 1981;72 (2):129–33.
  268. Bekoe EO, Kitcher C, Gyima NAM, Schwinger G, Frempong M. Medicinal Plants Used as Galactagogues. In: Pharmacognosy - Medicinal Plants [Internet]. IntechOpen; 2018 [cited 2023 Oct 12]. Available from: https://www.intechopen.com/chapters/64667
  269. Wuttke W, Jarry H, Christoffel V, Spengler B, Seidlová-Wuttke D. Chaste tree (Vitex agnus-castus)--pharmacology and clinical indications. Phytomedicine. 2003 May;10 (4):348–57.
  270. Prasad SK, Qureshi TN, Qureshi S. Mucuna pruriens seed powder feeding influences reproductive conditions and development in Japanese quail Coturnix coturnix japonica. Animal. 2009 Feb;3 (2):261–8.
  271. Jacobson H. Anti-Lactogenics – Herbs, Meds, Vitamins, and Foods that can undermine milk production - Hilary Jacobson [Internet]. 2022 [cited 2023 Nov 24]. Available from: https://hilaryjacobson.com/course_work/anti-lactogenics-herbs-meds-vitamins-and-foods-that-can-undermine-milk-production/, https://hilaryjacobson.com/course_work/anti-lactogenics-herbs-meds-vitamins-and-foods-that-can-undermine-milk-production/
  272. Mahdi AA, Shukla KK, Ahmad MK, Rajender S, Shankhwar SN, Singh V, et al. Withania somnifera Improves Semen Quality in Stress-Related Male Fertility. Evid Based Complement Alternat Med. 2009 Sep 29;
  273. Brandão Néto J, de Mendonça BB, Shuhama T, Marchini JS, Madureira G, Pimenta WP, et al. Zinc: an inhibitor of prolactin (PRL) secretion in humans. Horm Metab Res. 1989 Apr;21 (4):203–6.

 

Familial Hypercholesterolemia

ABSTRACT

 

Familial hypercholesterolemia (FH) is a prevalent, autosomal co-dominant disorder of lipid metabolism that results in elevated low-density lipoprotein cholesterol (LDL-C) levels and premature atherosclerosis. Screening for and identifying heterozygous FH in childhood is critical, given its high prevalence and asymptomatic presentation. Furthermore, treatment of FH in childhood is effective at lowering LDL-C levels and has the potential to reduce atherosclerotic cardiovascular disease (ASCVD) events in adulthood. Selective screening based on family history had previously been recommended to identify children and adolescents with FH or other lipid disorders. However, studies indicated that many individuals with heterozygous FH were missed with this approach, and therefore in 2011 the National Heart, Lung, and Blood Institute Expert Panel recommended universal screening of children and adolescents between ages 9 and 11 years and again at ages 17 to 21 years, in addition to selective screening, in order to identify pediatric individuals with heterozygous FH. This approach was affirmed in the 2018 American College of Cardiology and the American Heart Association (ACC/AHA) Cholesterol Guidelines, along with endorsing cascade screening as another reasonable approach to identifying children with FH. Once FH is diagnosed, prompt treatment with lifestyle modification should be initiated. When lifestyle interventions are not sufficient, pharmacotherapy using statins has been shown to be effective at lowering LDL-C, generally safe in short and medium- term studies, and may be beneficial at reducing ASCVD events. Other medications can be useful at lowering LDL-C in conjunction with statin therapy, although generally statins are sufficient in young patients. Homozygous FH is a rare disorder manifesting as extremely high LDL-C and ASCVD in childhood, requiring aggressive multimodal management. Overall, studies are needed to determine the optimal timing and intensity of statin therapy, and to better understand long-term safety and ASCVD outcomes in adulthood for lipid-lowering pharmacotherapy initiated in pediatric patients with heterozygous FH.

 

INTRODUCTION

 

Familial hypercholesterolemia (FH), as classically described, is the most common single gene disorder of lipoprotein metabolism and causes severely elevated low-density lipoprotein cholesterol (LDL-C) levels. The prevalence of FH is 1 in 200 to 1 in 300 individuals of different ethnicities (1,2), and it is strongly associated with premature coronary artery disease (CAD) (3). Data from observational studies suggest that untreated FH is associated with ~90-fold increase in mortality due to atherosclerotic cardiovascular disease (ASCVD) in young adults (4). Since early treatment may significantly reduce CAD-related morbidity and mortality in individuals with heterozygous FH (5), early identification and intervention during childhood may greatly improve outcomes in adulthood.

 

EPIDEMIOLOGY

 

The prevalence of FH varies substantially, depending upon the criteria used to define the disorder and the ancestry of the population.  Previously, the prevalence had been described as 1 in 500 individuals based on early work by Drs. Brown and Goldstein (6,7).  More recent analysis of white European populations, which tend to be less ethnically and racially diverse than the US, show higher prevalence rates (8,9).  An analysis using a modified version of the Dutch Lipid Clinic (DLC) criteria applied to participants in the 1999 to 2012 National Health and Education National Surveys (NHANES), a nationally representative survey of the US population, suggest FH affects 1 in 250 US adults (1). The prevalence of FH in US children and adolescents is not as well characterized, although presumably it is similar. Some estimate that as few as 10% of individuals with FH have been identified in the US.

 

PATHOPHYSIOLOGY AND GENETICS

 

FH was initially defined by Brown and Goldstein as a disorder or defect in the LDL receptor (LDL-R) (3). More recently, the description of FH has been expanded and used to describe any defects in LDL-C processing and/or signaling that may lead to a phenotype characteristic of FH (10). FH may be more common and complicated than previously thought, with many different genetic variants leading to pathogenesis. Overall, nine genes are causative for autosomal forms of FH, and up to 50 polymorphic loci contribute to polygenic susceptibility to elevated LDL- cholesterol levels (11).  Some etiologies for the FH phenotype include defects in apoB100 lipoprotein, the major atherogenic lipoprotein component of LDL-C, as well as gain of function mutations in proprotein convertase subtilisin/kexin type 9 (PCSK9), which promotes the degradation of the LDLR, resulting in reduced LDL-C clearance (7). Gene mutations in LDL-R, the apolipoprotein B gene, or in PCSK9 occur in approximately 93, 5, and 2 percent of individuals with a phenotype consistent with FH, respectively (12). The associated impairment in function of these receptors or proteins results in overall reduced clearance of LDL particles from the circulation and elevation in plasma LDL-C. There is also increased uptake of modified LDL by macrophage scavenger receptors, resulting in lipid accumulation in macrophages and foam cell formation, a precursor to atherosclerotic plaque development (13). Thus, the typical lipid profile of FH is characterized by elevated LDL-C (as high as 300 mg/dL) with subsequently increased total cholesterol (TC) levels; in general, triglycerides are normal, and high-density lipoprotein (HDL-C) can be low or normal. FH has an autosomal dominant inheritance pattern which results in hypercholesterolemia and early ASCVD events.

 

 

There are various genetic mutations that can cause FH, which can either be monogenic or polygenic in nature. See Table 1 below for a list of genes that are associated with FH.

 

Table 1. Genes Associated with FH

Monogenic

Autosomal Dominant

LDLR

APOB

PCSK9

APOE*

Autosomal Recessive

LDLRAP1

LIPA (Lysosomal acid lipase deficiency**)

ABCG5 and ABCG8 (Sitosterolemia**)

Polygenic

Genetic variants associated with FH which together can impact LDL-C levels

*Specific mutations or candidate regions associated with FH

**Refer to specific Endotext sections on lysosomal acid lipase deficiency and sitosterolemia for more information

 

In general, homozygotes for mutations in the LDL-R gene are more adversely affected than heterozygotes, reflecting a “gene dosing effect” of inheritance.  Unless there is consanguinity in a family in which heterozygous FH is present, homozygous FH is less prevalent and may affect as many as 1 in 160,000–300,000 individuals (14).  Additionally, the homozygous FH phenotype can be seen in compound heterozygotes which can occur in offspring of unrelated parents due to a different disease-causing mutation on each allele. The severity of the FH phenotype does not necessarily depend upon the presence of true homozygosity or compound heterozygosity inheritance; rather it is determined by the degree of disturbance in LDL metabolism.  For additional information on the genetics as well as pathophysiology, refer to “Familial Hypercholesterolemia: Genes and Beyond” by Warden et al., www.endotext.org (15).

 

FH PHENOTYPE

 

Clinical Symptoms

 

HOMOZYGOUS FH

 

Due to the excessively high plasma LDL-C levels in homozygous FH, cholesterol deposits are common in the tendons (xanthomas) and eyelids (xanthelasmas), and generally appear by one year of age. Tendon xanthomata are most common in the Achilles tendons and dorsum of the hands but can occur at other sites. Tuberous xanthomata typically occur over extensor surfaces such as the knee and elbow.  Planar xanthomas may occur on the palms of the hands and soles of the feet and are often painful.  Xanthelasmas are cholesterol-filled, soft, yellow plaques that usually appear on the medial aspects of the eyelids. Corneal arcus is a white or grey ring around the cornea (16).

 

HETEROZYGOUS FH

 

In addition to increased serum cholesterol and risk for premature coronary artery disease (see below), patients with heterozygous FH may have tendon xanthomas and corneal arcus that appear after the age of 20 years (16).

 

LDL-C Levels

 

In clinical practice, there is not a universal LDL-C threshold that determines a diagnosis of FH. Generally, the level of LDL-C that warrants further evaluation depends upon the age of the patient and whether additional family members have known hypercholesterolemia and/or early ASCVD. As suggested by recent guidelines (7,17), children and adolescents with a negative or unknown family history and LDL-C level persistently ≥190 mg/dL (4.9 mmol/L) suggests FH; in patients with a positive family history of hypercholesterolemia or early ASCVD, an LDL-C level of ≥160 mg/dL(4.1 mmol/L) is consistent with FH (see Table 2).

 

Table 2. Acceptable, Borderline, and High Lipid Levels for Children and Adolescents

Lipid

Low (mg/dl)

Acceptable (mg/dl

Borderline-High (mg/dl)

High (mg/dl)

TC

 

<170

170-199

>200

LDL-C

 

<110

110-129

>130

Non-HDL-C

 

<120

120-144

>145

Triglycerides

0-9 years

10-19 years

 

 

 

 

<75

<90

 

75-99

90-129

 

>100

>130

HDL-C

< 40

>45

 

 

Adapted from the NCEP expert panel on cholesterol levels in children (14)

 

Cardiovascular Disease

 

Cardiovascular risk in FH patients is determined by both the LDL-C concentration and by other traditional risk factors. Homozygotes have early-onset atherosclerosis, including myocardial infarction, in the first decade of life (reported as early as age two years), and are at increased risk for CAD-related mortality in the first and second decades (16).  Additionally, patients with homozygous FH can develop cholesterol and calcium deposits that can lead to aortic stenosis and occasionally to mitral regurgitation.

 

Heterozygotes are also at increased risk for early-onset CAD between the ages of 30-60 years (18). Children with heterozygous FH have thicker carotid intima-media thickness (cIMT), an anatomic measure of arterial thickness associated with atherosclerosis, compared to unaffected siblings and healthy controls (19,20). One study showed those treated with statin medications (HMG-CoA reductase inhibitors) at younger ages had less carotid atherosclerosis compared to the placebo group. Results from long-term studies of statins in children with FH are just emerging, and indicate that statin treatment during childhood may slow progression of cIMT and reduce the risk of CVD in adulthood (21).

 

SCREENING

 

Given the high prevalence of FH and the improved outcomes with early treatment, pediatric lipid screening has become very important for the detection of FH. However, the approach to lipid screening in childhood and adolescences has varied over the past decades and is somewhat controversial, with the US Preventative Services Task Force (USPSTF) recently concluding that that the current evidence is insufficient to assess the balance of benefits and harms of screening for lipid disorders in children and adolescents (22). Selective screening of young individuals with a family history of hypercholesterolemia and/or early CV events or patients at risk for atherosclerosis for other medical conditions has been recommended for several decades (23–25). However, screening individuals based only on family history may miss 30-50% of children with elevated LDL (24–26). Thus, the 2011 Expert Panel, supported by the more recent 2018 ACC/AHA Cholesterol Guidelines (27), recommended universal lipid screening, which involves screening in childhood at two time points, once between ages 9 and 11 years, and then again between ages 17 and 21 years (28). Universal screening is recommended in those not already selectively screened based on family history or personal risk factors (Table 3).

 

Selective screening for FH involves obtaining a fasting or non-fasting lipid panel in individuals ages 2 to 21 years with:

  1. Family history of early atherosclerosis or high cholesterol.
  2. Relatives of individuals with identified FH.

 

Lipid testing should also be performed in the presence of risk factors or medical diagnoses that increase risk for CVD (including hypertension, current cigarette smoking, body mass index ≥ 85th percentile, diabetes mellitus type I and II, chronic kidney disease/end-stage renal disease, chronic inflammatory diseases, human immunodeficiency virus infection, and nephrotic syndrome) (28).

 

Universal screening involves obtaining either a fasting lipid profile or a non-fasting non-HDL, (calculated by subtracting HDL from TC) in childhood at two time points:

  1. Between ages 9-11 years.
  2. Between ages 17-21 years.

 

Table 3. Screening for Hypercholesterolemia

Approach

Age in Years

Population

Selective

2-21

Family history of early atherosclerosis or high cholesterol

Presence of risk factors or medical conditions that increased early CVD risk*

Universal

9-11 and 17-21

All

*Selective screening is indicated in individuals with hypertension, current cigarette smoking, body mass index ≥ 85thpercentile, diabetes mellitus type I and II, chronic kidney disease/end-stage renal disease, chronic inflammatory diseases, human immunodeficiency virus infection, and nephrotic syndrome

 

DIAGNOSIS

 

The diagnosis of FH can be made clinically and through genetic testing; genotype needs to be interpreted in the context of phenotype. For heterozygous FH, the clinical diagnosis is made based on the presence of high levels of total and LDL cholesterol in combination with one or more of following (17):

 

  1. Family history of hypercholesterolemia (especially in children) or known FH.
  2. History of premature CAD in the patient or in family members.
  3. Physical examination findings of abnormal deposition of cholesterol in extravascular tissues (e.g., tendon xanthoma), although these rarely occur in childhood.

 

There are several clinical scoring systems used to diagnose FH, and these vary based on the weight given for each diagnostic criteria (11).  In general, clinical diagnosis of homozygous FH can be made in individuals with the following criteria (14):

 

  1. Untreated LDL-C >500mg/dL (>13 mmol/L) or treated LDL-C ≥300 mg/dL (>8 mmol/L), AND
  2. Cutaneous or tendon xanthoma before age 10 years, OR
  3. Elevated LDL-C levels consistent with heterozygous FH in both parents.

 

There are various disease states or other factors that can increase LDL-C levels and should be considered when diagnosing FH.  Some of these include the following:

 

  • Obesity
  • Hypothyroidism
  • Diabetes mellitus
  • Nephrotic syndrome
  • Chronic renal failure
  • Cholestasis
  • Biliary atresia
  • Hepatitis
  • Biliary cirrhosis
  • HIV infection/AIDS
  • Various drugs/medications
  • Alcohol
  • Pregnancy
  • Very low carbohydrate ketogenic diets

 

Genetic Testing

 

Identifiable gene defects in LDLR, APOB, or PCSK9 have been identified in 60 to 80% of individuals with a heterozygote FH phenotype. Genetic testing has not routinely been performed in the clinical setting due to concerns about cost to the patient and because it was not likely to alter management, given that treatment decisions were usually based on LDL-C levels. However, FH genetic testing has recently been recommended to become the standard of care for patients with definite or probable FH, and the rationale for such testing includes the following: 1) facilitation of definitive diagnosis of FH lowers the concern for other secondary causes of high LDL-C; 2) pathogenic variants correlate with higher cardiovascular risk, which indicates the potential need for more aggressive lipid lowering; 3) increase in initiation of and adherence to therapy; and 4) cascade testing of at-risk relatives (29). Overall, the clinical significance of normal or moderately elevated LDL-C levels in the setting of a genetic defect in the LDLR or other possibly pathogenic defects is unknown.

 

TREATMENT

 

The guidelines for initiating treatment in patients with the FH phenotype are based on age, severity of LDL-C elevation, as well as family and medical histories. Lifestyle therapy is recommended for all children and adolescents with LDL-C levels ≥ 130 mg/dL. If lifestyle intervention is insufficient, medications can be considered in children beginning at age 10 years, or as early as age 8 in high-risk patients and in the presence of a very high-risk family history. For healthy children and adolescents ages 10-21 years, lifestyle therapy should be provided to those with an LDL-C ≥ 130 mg/dL, and medication should be initiated if LDL-C remains ≥ 190 mg/dL despite 6 or more months of lifestyle modification. If there is a family history of early atherosclerotic disease, then medication should be started in individuals with LDL-C levels ≥ 160 mg/dL who do not respond sufficiently to lifestyle modification. If an individual has a high-risk medical condition, as noted above, medication can be considered for those with a persistently elevated LDL-C ≥ 130 mg/dL. In general, the goal of treatment is to maintain an LDL-C level ≤ 130 mg/dL or ≥ 50% reduction in LDL concentration; lower ranges may be considered in high-risk patients. Medications should be initiated in all patients with homozygous FH at the time of diagnosis, regardless of age, and additional treatments should also be considered.

 

Lifestyle Treatment

 

The mainstay of treatment for pediatric lipid disorders is lifestyle modification. A low saturated fat diet, without trans-fat, and high in fruits and vegetables, is the recommended diet for lowering LDL-C. This dietary approach has been shown to be both safe and beneficial in the general pediatric population (28,30,31).  Additionally, nutritional and physical activity interventions have been shown to lower LDL-C and improve CVD risk factors in children with obesity in meta-analyses (32). Despite this, in adults with FH, lifestyle modifications have been shown to only lower LDL-C modestly (33).  Furthermore, the effect of physical activity on LDL-C levels has not been well studied in children with FH.

 

Pharmacotherapy

 

STATINS  

 

The majority of patients with FH are treated with medications, and statins are the recommended first line pharmacotherapy. In a Cochrane meta-analysis of pediatric patients with FH, statins were shown to lower LDL-C by 32% (34). Furthermore, more intensive statin therapy in high doses has been shown to lower LDL-C even more significantly, by up to 50% (35,36).  Follow-up data from a statin trial in pediatric Dutch patients suggest efficacy and safety, as well as decreased atherosclerosis compared to the subjects’ parents (37). Most recently, the use of statins in children with FH have shown reduced CVD risk in adulthood for these patients compared to their untreated parents with FH, after a 20 year follow-up (21). Although this study was not a controlled or placebo study, it suggests that long-term statin use in childhood may prevent ASCVD compared to not treating.

 

Several different formulations of statin therapy are available and approved by the US Food and Drug Administration (FDA) for use in children. Treatment is initiated at a low dose (generally 5-20mg depending on the statin potency), which is given once a day, often at night. If needed, the dose is increased to meet the goals of therapy. Side effects with statins are rare, but include myopathy, new-onset type 2 diabetes mellitus (reported in adult primary prevention statin trials), and hepatic enzyme elevation. In pediatric clinical trials, rates of side effects with statin therapy were low and adherence to statin therapy was generally good (38). Side effects of statins are more likely at higher doses and in patients taking other medications, particularly cyclosporine, azole antifungal agents, and other medications and foods (such as grapefruit) that impact the cytochrome P450 system. Adolescent females should be counseled about the possibility of drug teratogenicity and appropriate contraceptive methods while receiving statin therapy. Additionally, providers should be aware that oral contraceptive pills can increase lipid levels.

 

The NHLBI guidelines recommend the following baseline laboratory evaluation when initiating statin therapy (28):

 

  • Fasting lipid profile.
  • Serum creatine kinase (CK).
  • Hepatic enzymes (i.e., serum alanine aminotransferase [ALT] and aspartate aminotransferase [AST]).

 

Screening for type 2 diabetes is also reasonable prior to starting statins. Fasting lipid profiles are repeated at four weeks after the initiation of statin therapy to titrate dose and are repeated every six months in patients on stable therapy. Liver function tests, CK, and hemoglobin A1C should be obtained if signs of adverse effects arise, and may be obtained at regular intervals, for example after dose changes based on best clinical judgement. Ongoing monitoring of growth, other measures of general and cardiovascular health, and review for the presence of additional ASCVD risk factors, such as smoking exposure, should also occur at each visit.

 

BILE ACID BINDING RESINS

 

Although bile acid binding resins or bile acid sequestrants have been shown to lower LDL-C by ~10-20% in pediatric trials (39,40), they are often difficult to tolerate given their unpalatability and associated adverse effects (such as bloating and constipation) (41). For these reasons, bile acid binding resins are used relatively infrequently. However, they may be useful in combination with a statin for patients who fail to meet target LDL-C levels (42). The sequestrants are not absorbed systemically, remain in the intestines, and are excreted along with bile containing cholesterol. Therefore, they are considered to be very safe. They can be used in patients who prefer to avoid statins, although they may not lower LDL-C sufficiently to achieve goal levels.

 

CHOLESTEROL ABSORPTION INHIBITORS (EZETIMIBE)

 

Ezetimibe is a lipid-lowering mediation that inhibits absorption of cholesterol and plant sterols in the intestines. This agent can be useful in pediatric patients with FH who are not able to reach LDL-C treatment goals on high-intensity statin therapy. Ezetimibe further lowers serum LDL-C and in adults has been shown to improve cardiovascular outcomes without altering the side effect profile (43–45). Specifically in the pediatric population, ezetimibe has been shown to be safe and effective at lowering LDL-C by up to almost 30%, even when used as monotherapy (44,46,47)

 

PCSK9 INHIBITORS

 

PCSK9 inhibitors (evolocumab and alirocumab) are human monoclonal antibodies that bind to PCSK9 and promote plasma LDL cholesterol clearance. In Europe, evolocumab is approved in adolescents (≥12 years old) with homozygous FH. In the US, alirocumab is approved only for use in adult patients, and evolocumab is approved for use in adults with heterozygous FH and in homozygous FH, ages 13 and older, who have not responded to other LDL-C lowering therapies. Overall, PCSK9 inhibitors appear to have a good safety and side effect profile in adults (48).  These medications have been shown to be very effective, reducing LDL-C by more than 15% in patients with homozygous FH and by 35% in patients with heterozygous FH (49–51). The main disadvantage of PCSK9 inhibitors is that they require injection for administration; cost is also a concern.

 

Inclisiran is another medical therapy that targets PCSK9 synthesis through a different mechanism.  It is a small interfering RNA molecule that triggers the breakdown of messenger RNA coding for the PCSK9 protein. This medication has been recently approved for clinical use and has been shown to be safe and effective in adult patients with heterozygous FH (49). Clinical trials of inclisiran are currently underway in adolescent patients with heterozygous and homozygous FH (50).

 

EMERGING MEDICAL THERAPIES

 

There are several promising therapies that aim to reduce LDL-C through different approaches.  These include Bempedoic Acid, Lomitapide, and Evinacumab.  Bempedoic acid blocks the cholesterol biosynthetic pathway upstream of HMG-CoA reductase through inhibition of adenosine triphosphate citrate lyase. This therapeutic agent has been shown to be effective in treating statin-resistant hypercholesterolemia and in reducing ASCVD in adults (51,52). Bempedoic acid is currently being studied in children with heterozygous FH.

 

Microsomal triglyceride transfer protein (MTP) plays an essential role in the formation of apoB-containing lipoproteins and has been shown to be inhibited by Lomitapide (53). This drug can reduce LDL-C by approximately 58% and has been approved for use in adults with homozygous FH.  So far, Lomitapide has been observed to be safe and effective in pediatric patients with homozygous FH (54).

 

Evinacumab is a human monoclonal antibody that targets angiopoietin-like 3 (ANGPTL3), which results in the reduction of LDL-C levels via an LDL-receptor independent mechanism (55–57).  This drug has been shown to significantly reduce LDL-C in patients with homozygous FH who show little to no LDL-receptor activity and who have had poor response to other treatments (58). Additionally, patients with heterozygous FH have also shown improvements in LDL-C by 50% reduction (59).  Evinacumab was approved in early 2021 by the FDA for the treatment of homozygous FH patients starting at age 12 years (60).

 

Lipoprotein Apheresis

 

Although statin therapy has been shown to be effective in reducing LDL and prolonging life expectancy in patients with homozygous FH (61), medical treatment alone may not be adequate to achieve recommended treatment goals. Therefore, it is suggested that lipoprotein apheresis (LA) be initiated in patients with homozygous FH as young as 2 years. The efficacy is dependent upon the type of apheresis but can reduce LDL-C by as much as 45 to 80%. Studies in pediatric patients are limited, but there is some evidence suggesting that LA therapy is safe and effective in children with homozygous FH (62).

 

SUMMARY

 

FH is an autosomal dominant disorder of LDL metabolism that affects 1 in 200 to 300 individuals. Screening involving lipid measurements, family and medical history, and physical examination is needed to identify affected individuals; cascade screening can be helpful. Lifestyle modification is the first-line therapy for hyperlipidemia in pediatric patients but is usually not sufficient to achieve goal LDL levels. Available evidence suggests that treatment with lipid-lowering pharmacotherapy, such as statins, is effective and generally safe in the short and medium-term.  However, further studies are needed to determine the long-term safety and efficacy in preventing ASCVD of lipid lowering medication in pediatric patients with FH.

 

REFERENCES

 

  1. de Ferranti SD, Rodday AM, Mendelson MM, Wong JB, Leslie LK, Sheldrick RC. Prevalence of Familial Hypercholesterolemia in the 1999 to 2012 United States National Health and Nutrition Examination Surveys (NHANES). Circulation. 2016 Mar 15;133(11):1067-72.
  2. Toft-Nielsen F, Emanuelsson F, Benn M. Familial Hypercholesterolemia Prevalence Among Ethnicities-Systematic Review and Meta-Analysis. Front Genet. 2022 Feb 3;13:840797.
  3. 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 Internet. Published online July 1973. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=302426&tool=pmcentrez&rendertype=abstract
  4. Neil A, Cooper J, Betteridge J, Capps N, McDowell I, Durrington P. Reductions in all-cause, cancer, and coronary mortality in statin-treated patients with heterozygous familial hypercholesterolaemia: a prospective registry study. Eur Heart J Internet. Published online 2008. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2577142&tool=pmcentrez&rendertype=abstract
  5. Versmissen J, Oosterveer DM, Yazdanpanah M, Defesche JC, Basart DCG, Liem AH. Efficacy of statins in familial hypercholesterolaemia: a long term cohort study. BMJ Internet. Published online 2008. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2583391&tool=pmcentrez&rendertype=abstract
  6. Brown MS, Goldstein JL. Expression of the familial hypercholesterolemia gene in heterozygotes: mechanism for a dominant disorder in man. Sci Internet. Published online July 5, 1974. http://www.ncbi.nlm.nih.gov/pubmed/4366052
  7. Goldberg AC, Hopkins PN, Toth PP, Ballantyne CM, Rader DJ, Robinson JG. 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 Internet. Published online 2011. http://www.ncbi.nlm.nih.gov/pubmed/21600525
  8. Watts GF, Shaw JE, Pang J, Magliano DJ, Jennings GLR, Carrington MJ. Prevalence and treatment of familial hypercholesterolaemia in Australian communities. Int J Cardiol Internet. Published online April 15, 2015. http://www.ncbi.nlm.nih.gov/pubmed/25791093
  9. Austin MA, Hutter CM, Zimmern RL, Humphries SE. Familial hypercholesterolemia and coronary heart disease: a HuGE association review. Am J Epidemiol Internet. Published online September 1, 2004. http://www.ncbi.nlm.nih.gov/pubmed/15321838
  10. 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 Internet. Published online 2011. http://www.ncbi.nlm.nih.gov/pubmed/21600530
  11. Berberich AJ, Hegele RA. The complex molecular genetics of familial hypercholesterolaemia. Nat Rev Cardiol Internet. 2019;16(1):9-20. doi:10.1038/s41569-018-0052-6
  12. Talmud PJ, Shah S, Whittall R, Futema M, Howard P, Cooper JA. Use of low-density lipoprotein cholesterol gene score to distinguish patients with polygenic and monogenic familial hypercholesterolaemia: a case-control study. Lancet. http://www.ncbi.nlm.nih.gov/pubmed/23433573
  13. Epstein FH, Steinberg D, Parthasarathy S, Carew TE, Khoo JC, Witztum JL. Beyond Cholesterol. N Engl J Med Internet. Published online 1989. http://www.ncbi.nlm.nih.gov/pubmed/2648148
  14. Cuchel M, Bruckert E, Ginsberg HN, Raal FJ, Santos RD, Hegele RA. 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 Internet. Published online August 21, 2014. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4139706&tool=pmcentrez&rendertype=abstract
  15. Warden BA, Fazio S, Shapiro MD, et al. Familial Hypercholesterolemia: Genes and Beyond. editors., ed. Endotext Internet South Dartm MA MDTextcom Inc. Published online 2018.
  16. Naoumova RP, Thompson GR, Soutar AK. Current management of severe homozygous hypercholesterolaemias. Curr Opin Lipidol Internet. Published online 2004. http://www.ncbi.nlm.nih.gov/pubmed/15243214
  17. Nordestgaard BG, Chapman MJ, Humphries SE, Ginsberg HN, Masana L, Descamps OS. 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 Internet. Published online 2013. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3844152&tool=pmcentrez&rendertype=abstract
  18. Ose L. Familial hypercholesterolemia from children to adults. http://www.ncbi.nlm.nih.gov/pubmed/12652097
  19. Wiegman A, Groot E, Hutten BA, Rodenburg J, Gort J, Bakker HD. Arterial intima-media thickness in children heterozygous for familial hypercholesterolaemia. Lancet. Internet. Published online January 31, 2004. http://www.ncbi.nlm.nih.gov/pubmed/15070569
  20. Tonstad S, Joakimsen O, Stensland-Bugge E, Leren TP, Ose L, Russell D. Risk factors related to carotid intima-media thickness and plaque in children with familial hypercholesterolemia and control subjects. Arter Thromb Vasc Biol Internet. Published online 1996. http://www.ncbi.nlm.nih.gov/pubmed/8696963
  21. Kusters DM, Ph D, Hof MH, Ph D. 20-Year Follow-up of Statins in Children with Familial Hypercholesterolemia. :2019 1547-56.
  22. Association AM. Lipid Screening in Childhood and Adolescence for Detection of Familial Hypercholesterolemia. Evid Rep Syst Rev US Prev Serv Task Force. 2019;98101(6):645-655.
  23. Daniels G SR, F.R. Lipid screening and cardiovascular health in childhood. Pediatr Internet. Published online 2008. http://www.ncbi.nlm.nih.gov/pubmed/18596007
  24. Executive Summary of The Third Report of The National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, And Treatment of High Blood Cholesterol In Adults (Adult Treatment Panel III. JAMA Internet. Published online May 16, 2001. http://www.ncbi.nlm.nih.gov/pubmed/11368702
  25. National Cholesterol Education Program (NCEP): highlights of the report of the Expert Panel on Blood Cholesterol Levels in Children and Adolescents. Pediatr Internet. Published online 1992. http://www.ncbi.nlm.nih.gov/pubmed/1741227
  26. Dennison BA, Jenkins PL, Pearson TA. Challenges to implementing the current pediatric cholesterol screening guidelines into practice. Pediatr Internet. Published online 1994. http://www.ncbi.nlm.nih.gov/pubmed/8065853
  27. Grundy SM, Stone NJ, Bailey AL, Jones DW, Beam C, Lloyd-jones D. In: 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 WRIT. ; 2018.
  28. Expert panel on integrated guidelines for cardiovascular health and risk reduction in children and adolescents: summary report. Pediatr Internet. Published online 2011. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4536582&tool=pmcentrez&rendertype=abstract
  29. Sturm AC, Knowles JW, Gidding SS, et al. Clinical Genetic Testing for Familial Hypercholesterolemia JACC Scienti fi c Expert Panel. Lipid Screen Child. 2019;1(10):1437-1438.
  30. Niinikoski H, Lagström H, Jokinen E, Siltala M, Rönnemaa T, Viikari J. Impact of repeated dietary counseling between infancy and 14 years of age on dietary intakes and serum lipids and lipoproteins: the STRIP study. Circ Internet. Published online August 28, 2007. http://www.ncbi.nlm.nih.gov/pubmed/17698729
  31. Efficacy and safety of lowering dietary intake of fat and cholesterol in children with elevated low-density lipoprotein cholesterol. The Dietary Intervention Study in Children (DISC). The Writing Group for the DISC Collaborative Research Group. JAMA Internet. Published online May 10, 1995. http://www.ncbi.nlm.nih.gov/pubmed/7723156
  32. Ho M, Garnett SP, Baur L, Burrows T, Stewart L, Neve M. Effectiveness of lifestyle interventions in child obesity: systematic review with meta-analysis. Pediatr Internet. Published online 2012. http://www.ncbi.nlm.nih.gov/pubmed/23166346
  33. Pt K, Se H. Statins for children with familial hypercholesterolemia (Review. Published online 2019.
  34. Vuorio A, Kuoppala J, Kovanen PT, Humphries SE, Tonstad S, Wiegman A. Statins for children with familial hypercholesterolemia. Cochrane Database Syst Rev Internet. Published online 2014. http://www.ncbi.nlm.nih.gov/pubmed/25054950
  35. McCrindle BW, Ose L, Marais AD. Efficacy and safety of atorvastatin in children and adolescents with familial hypercholesterolemia or severe hyperlipidemia: a multicenter, randomized, placebo-controlled trial. J Pediatr. 2003;143(1):74-80. doi:10.1016/S0022-3476(03)00186-0
  36. Avis HJ, Hutten BA, Gagné C, et al. Efficacy and safety of rosuvastatin therapy for children with familial hypercholesterolemia. J Am Coll Cardiol. 2010;55(11):1121-1126. doi:10.1016/j.jacc.2009.10.042
  37. Kusters DM, Avis HJ, Groot E, Wijburg FA, Kastelein JJP, Wiegman A. Ten-year follow-up after initiation of statin therapy in children with familial hypercholesterolemia. JAMA Internet. Published online September 10, 2014. http://www.ncbi.nlm.nih.gov/pubmed/25203086
  38. Braamskamp MJAM, Kusters DM, Avis HJ, Smets EMA, Wijburg FA, Kastelein JJP. Long-term statin treatment in children with familial hypercholesterolemia: more insight into tolerability and adherence. Paediatr Drugs Internet. Published online 2015. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4372689&tool=pmcentrez&rendertype=abstract
  39. Tonstad S, Knudtzon J, Sivertsen M, Refsum H, Ose L. Efficacy and safety of cholestyramine therapy in peripubertal and prepubertal children with familial hypercholesterolemia. J Pediatr Internet. Published online July 1996. http://www.ncbi.nlm.nih.gov/pubmed/8757561
  40. Liacouras CA, Coates PM, Gallagher PR, Cortner JA. Use of cholestyramine in the treatment of children with familial combined hyperlipidemia. J Pediatr Internet. Published online 1993. http://www.ncbi.nlm.nih.gov/pubmed/8441109
  41. McCrindle BW, O’Neill MB, Cullen-Dean G, Helden E. Acceptability and compliance with two forms of cholestyramine in the treatment of hypercholesterolemia in children: a randomized, crossover trial. J Pediatr Internet. Published online 1997. http://www.ncbi.nlm.nih.gov/pubmed/9042130
  42. McCrindle BW, Helden E, Cullen-Dean G, Conner WT. A randomized crossover trial of combination pharmacologic therapy in children with familial hyperlipidemia. http://www.ncbi.nlm.nih.gov/pubmed/12032266
  43. Ip C, Jin D, Gao J, Meng Z, Meng J, Tan Z. Effects of add-on lipid-modifying therapy on top of background statin treatment on major cardiovascular events: A meta-analysis of randomized controlled trials. Int J Cardiol Internet. Published online July 15, 2015. http://www.ncbi.nlm.nih.gov/pubmed/25965621
  44. Kusters DM, Caceres M, Coll M, Cuffie C, Gagné C, Jacobson MS. Efficacy and safety of ezetimibe monotherapy in children with heterozygous familial or nonfamilial hypercholesterolemia. J Pediatr Internet. Published online 2015. http://www.ncbi.nlm.nih.gov/pubmed/25841542
  45. van der Graaf A, Cuffie-Jackson C, Vissers MN, Trip MD, Gagné C, Shi G, Veltri E, Avis HJ, Kastelein JJ. Efficacy and safety of coadministration of ezetimibe and simvastatin in adolescents with heterozygous familial hypercholesterolemia. J Am Coll Cardiol. 2008 Oct 21;52(17):1421-9
  46. Yeste D, Chacón P, Clemente M, Albisu MA, Gussinyé M, Carrascosa A. Ezetimibe as monotherapy in the treatment of hypercholesterolemia in children and adolescents. J Pediatr Endocrinol Metab JPEM. 2009;22(6):487-492. doi:10.1515/jpem.2009.22.6.487
  47. Clauss S, Wai KM, Kavey REW, Kuehl K. Ezetimibe treatment of pediatric patients with hypercholesterolemia. J Pediatr. 2009;154(6):869-872. doi:10.1016/j.jpeds.2008.12.044
  48. Zhang XL, Zhu QQ, Zhu L, Chen JZ, Chen QH, Li GN. Safety and efficacy of anti-PCSK9 antibodies: a meta-analysis of 25 randomized, controlled trials. BMC Med Internet. Published online 2015. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4477483&tool=pmcentrez&rendertype=abstract
  49. Raal FJ, Kallend D, Ray KK, et al. Inclisiran for the Treatment of Heterozygous Familial Hypercholesterolemia. N Engl J Med. 2020;382(16):1520-1530. doi:10.1056/NEJMoa1913805
  50. Reijman MD, Schweizer A, Peterson ALH, et al. Rationale and design of two trials assessing the efficacy, safety, and tolerability of inclisiran in adolescents with homozygous and heterozygous familial hypercholesterolaemia. Eur J Prev Cardiol. 2022;29(9):1361-1368. doi:10.1093/eurjpc/zwac025
  51. Goldberg AC, Leiter LA, Stroes ESG, et al. Effect of Bempedoic Acid vs Placebo Added to Maximally Tolerated Statins on Low-Density Lipoprotein Cholesterol in Patients at High Risk for Cardiovascular Disease: The CLEAR Wisdom Randomized Clinical Trial. JAMA. 2019;322(18):1780-1788. doi:10.1001/jama.2019.16585
  52. Nissen SE, Lincoff AM, Brennan D, et al. Bempedoic Acid and Cardiovascular Outcomes in Statin-Intolerant Patients. N Engl J Med. 2023;388(15):1353-1364. doi:10.1056/NEJMoa2215024
  53. Alonso R, Cuevas A, Mata P. Lomitapide: a review of its clinical use, efficacy, and tolerability. Core Evid. 2019;14:19-30. doi:10.2147/CE.S174169
  54. Ben-Omran T, Masana L, Kolovou G, et al. Real-World Outcomes with Lomitapide Use in Paediatric Patients with Homozygous Familial Hypercholesterolaemia. Adv Ther. 2019;36(7):1786-1811. doi:10.1007/s12325-019-00985-8
  55. Adam RC, Mintah IJ, Alexa-Braun CA, et al. Angiopoietin-like protein 3 governs LDL-cholesterol levels through endothelial lipase-dependent VLDL clearance. J Lipid Res. 2020;61(9):1271-1286. doi:10.1194/jlr.RA120000888
  56. Stitziel NO, Khera AV, Wang X, et al. ANGPTL3 Deficiency and Protection Against Coronary Artery Disease. J Am Coll Cardiol. 2017;69(16):2054-2063. doi:10.1016/j.jacc.2017.02.030
  57. Rhee JW, Wu JC. Dyslipidaemia: In vivo genome editing of ANGPTL3: a therapy for atherosclerosis? Nat Rev Cardiol. 2018;15(5):259-260. doi:10.1038/nrcardio.2018.38
  58. Raal FJ, Rosenson RS, Reeskamp LF, et al. Evinacumab for Homozygous Familial Hypercholesterolemia. N Engl J Med. 2020;383(8):711-720. doi:10.1056/NEJMoa2004215
  59. Rosenson RS, Burgess LJ, Ebenbichler CF, et al. Evinacumab in Patients with Refractory Hypercholesterolemia. N Engl J Med. 2020;383(24):2307-2319. doi:10.1056/NEJMoa2031049
  60. Markham A. Evinacumab: First Approval. Drugs. 2021;81(9):1101-1105. doi:10.1007/s40265-021-01516-y
  61. Raal FJ, Pilcher GJ, Panz VR, Deventer HE, Brice BC, Blom DJ. Reduction in mortality in subjects with homozygous familial hypercholesterolemia associated with advances in lipid-lowering therapy. Circulation. 2011;124(20):2202-2207.
  62. Luirink IK, Determeijer J, Hutten BA, Wiegman A, Bruckert E, Schmitt CP. Efficacy and safety of lipoprotein apheresis in children with homozygous familial hypercholesterolemia: A systematic review. J Clin Lipidol. 2019;13(1):31-39.

 

Use of Lipid Lowering Medications in Youth

ABSTRACT

 

The first comprehensive pediatric dyslipidemia guidelines were published by the National Cholesterol Education Program’s Expert Panel on Blood Cholesterol Levels in Children and Adolescents in 1992 and were updated by the American Academy of Pediatrics (AAP) in 1998. In 2008 the AAP issued an updated clinical report detailing recommendations for screening and evaluation of cholesterol levels in children and adolescents as well as prevention and treatment strategies. Since the publication of the first guidelines, rates of pediatric obesity have significantly increased, resulting in a concomitant increase in dyslipidemia. Recently, options for pharmacotherapeutic interventions in pediatric patients have expanded with new FDA approved indications of several lipid lowering medications, as well as additional safety and efficacy data. In 2011, The National Heart Lung and Blood Institute (NHLBI) published its comprehensive report Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents. As with previous guidelines, lifestyle modifications with an emphasis on a heart-healthy diet and daily moderate to vigorous exercise remain an integral part of treatment for pediatric lipid disorders; however, the recommendations for patients requiring management with pharmacotherapy have changed, and will be the focus of this discussion.

 

INTRODUCTION

 

The diagnosis, treatment, and monitoring of dyslipidemia in youth has undergone significant transformations in recent years.  As detailed by the Pathobiological Determinants of Atherosclerosis in Youth (PDAY) and Bogalusa Heart studies, dyslipidemia plays a vital role in both the initiation, as well as the progression of atherosclerotic lesions in children and adolescents (1-3). Because of their role in premature cardiovascular disease, control of dyslipidemia provides clinicians with an opportunity for reducing morbidity and mortality.  Observational data from individuals with genetic mutations that lower atherogenic cholesterol, low-density lipoprotein cholesterol (LDL-C) and non-high-density lipoprotein cholesterol (non-HDL-C), over a lifetime are associated with fewer events and longer life expectancy (4, 5). While these observations are very encouraging, it is not known if achieving the same level of lipid lowering with medications over decades will offer the same protective effects as observed in individuals with life-long lower cholesterol secondary to a genetic mutation (6). Table 1 provides a comparison across the evolution of guideline recommendations for the initiation of pharmacologic intervention with the goal of balancing risk and benefit (7-11). Table 2 details the risk factors and risk conditions described in the NHLBI guidelines (9).

 

Table 1. Comparison of Recommendations for Treatment

Guidelines

NCEP, AAP – 1992 & 1998

AAP – 2008

NHLBI, AAP – 2011

Pharmacologic Treatment Initiation Parameters*

Age > 10 years with LDL-C

Age ≥ 8 years with LDL-C

Ages 10-21 years with LDL-C

 

o

≥ 190 mg/dL

 

o

≥ 190 mg/dL

 

o

≥ 190 mg/dL

 

o

> 160 mg/dL in addition to a positive family history of premature CVD or presence of at least 2 CVD risk factors in the child/adolescent

 

o

≥ 160 mg/dL in addition to a positive family history of premature CVD or presence of risk factors

 

o

160-189 mg/dL in addition to a positive family history of premature CVD or presence of 1 high level risk factor/condition or presence of 2 moderate level risk factors/conditions

 

 

 

 

 

 

 

 

 

 

 

o

≥ 130 mg/dL in addition to presence of diabetes mellitus

 

 

 

 

   

 

 

 

 

 

 

   

Age < 8 years with LDL-C:

 

o

130-159 mg/dL in addition to the presence of 2 high level risk factors/conditions or 1 high level and at least 2 moderate level risk factors/conditions

 

 

   

 

o

≥ 500 mg/dL

 

 

 

   

 

   

 

 

 

 

   

 

   

 

 

 

 

   

 

   

Age < 10 years with severe hyperlipidemia or high-risk conditions associated with serious morbidity

 

 

   

 

   

 

 

 

   

 

   

 

*After an adequate trial of diet and lifestyle management.

 

   

 

   

Ages 8-9 years with LDL-C levels consistently ≥ 190 mg/dL in addition to a positive family history OR presence of risk factors

 

   

 

   

 

 

   

 

   

 

 

 

 

 

 

 

 

Pharmacologic Medication Recommendations

Bile acid sequestrants

Bile acid sequestrants

Statins

 

   

Cholesterol absorption inhibitors

 

 

 

 

 

 

Statins

 

 

 

 

 

Table 2. NHLBI Risk Factors and Risk Conditions (9)

NHLBI, AAP 2011 Guidelines: Risk Factors and Risk Conditions

High Level Risk Factors

•       Hypertension requiring drug therapy

•       Tobacco use

•       BMI ≥ 97th percentile

•       High risk conditions

Moderate Level Risk Factors

•       Hypertension not requiring drug therapy

•       BMI ≥ 95th percentile, < 97th percentile

•       HDL < 40 mg/dL

•       Moderate risk conditions

High Risk Conditions

•       T1DM and T2DM

•       CKD, ESRD, post-renal transplant

•       Post-orthotopic heart transplant

•       Kawasaki disease with current aneurysms

Moderate Risk Conditions

•       Kawasaki disease with regressed aneurysms

•       Chronic inflammatory disease

•       HIV infection

•       Nephrotic syndrome

 

PHARMACOTHERAPEUTIC TREATMENT IN YOUTH

 

The treatment of youth with lipid lowering medications presents some unique challenges and consideration due to their developmental stage, the possibility of extended durations of treatment, and the potential use of concurrent medications that may be counter-productive by increasing lipid levels. As patients progress into adolescence it is particularly important for the patient to understand not only the need for their lipid lowering therapies, but also the consequences of non-compliance. Counseling regarding pharmacotherapy should begin at an early age with developmentally appropriate explanations and expand as patients mature. During adolescence when patients are developing their independence counseling that addresses how to integrate their therapy into their own social norms is important for achieving compliance to both pharmacotherapy as well as lifestyle modifications. Additionally, the cost of therapy significantly impacts compliance and should be factored into therapy decisions especially as youth transition into adulthood and may be faced with changes in insurance coverage. It is also critical to continually readdress the correct use of medications as patients are likely to be on multiple therapies and adolescents will begin some their own medication management as they mature. Regular monitoring for adverse events and side effects of therapy is essential as youth will have a greater lifetime exposure compared to adults and long-term data is generally limited.

 

As with all pharmacotherapy careful consideration should be given to potential drug interactions including those that may increase lipid levels. It is not uncommon for adolescent patients to be prescribed medications which have the potential to negatively impact lipid levels such as systemic steroids or oral contraceptive pills. Each patient case must be evaluated on an individual basis to determine the risk and benefit of prescribing medications which negatively alter lipid levels for patients also utilizing lipid lowering therapies. It should be noted there is significant risk for adolescent females should they become pregnant while taking lipid lowering medications as some have demonstrated a negative impact on fetal development. Adolescent females should be counselled regarding pregnancy and methods of contraception should be discussed. The US Medical Eligibility Criteria for Contraceptive Use compiled by the CDC details several contraceptive methods where the benefits generally outweigh any theoretical or proven risk for patients with hyperlipidemias (12).

 

HMG-CoA REDUCTASE INHIBITORS

 

HMG-CoA reductase inhibitors, or statins, are recommended as first line treatment of youth with severe dyslipidemia who fail non-pharmacologic interventions (i.e., diet and lifestyle modification) (8-11). Statins first debuted in clinical practice in 1987 with the FDA’s approval of lovastatin. At present, there are seven HMG-CoA reductase inhibitors with FDA approval, at varying dosages, for youth with heterozygous familial hypercholesterolemia. Lovastatin, simvastatin, atorvastatin and fluvastatin are approved for children 10 years of age and older. Pitavastatin and pravastatin are approved starting at 8 years of age and rosuvastatin is indicated in children as early as age 6 (13-21). Table 3 provides a summary of HMG-CoA reductase inhibitors, pediatric approval and indications, recommended dosing ranges, comments on dosing, and supporting clinical trials (13-21).

 

Table 3. HMG-CoA Reductase Inhibitors

Medication

Pediatric Approvals & Indications

Dosing

Comments

Supporting

Clinical Trials

Atorvastatin

Age 10-17
Heterozygous familial hypercholesterolemia

10-20 mg/day

May be titrated at ≥ 4-week intervals

McCrindle, et al (22)

Fluvastatin

Age 10-16
Heterozygous familial hypercholesterolemia

20-80 mg/day

May be titrated at ≥ 6-week intervals

van der Graaf, et al (23)

Lovastatin

Age 10-17
Heterozygous familial hypercholesterolemia

10-40 mg/day

Initiated at 20 mg/day for ≥20% LDL reduction, may be titrated at ≥ 4-week intervals

Clauss, et al (24)
Lambert, et al (25)
Stein, et al (26)

Pravastatin

Age 8 and older
Heterozygous familial hypercholesterolemia

20-40 mg/day

Age 8-13: 20 mg/day
Age 14-18: 40 mg/day

Knipscheer, et al (27)
Wiegman, et al (28)
Rodenburg, et al (29)

Rosuvastatin

Age 6 and older
Heterozygous familial hypercholesterolemia

5-20 mg/day

May be titrated at ≥ 4-week intervals

Avis, et al (30)

Simvastatin

Age 10-17
Heterozygous familial hypercholesterolemia

10-40 mg/day

May be titrated at ≥ 4-week intervals

de Jongh, et al (31)
de Jongh, et al (32)

Pitavastatin

Age 8 and older
Heterozygous familial hypercholesterolemia

1-4 mg/day

May be titrated at ≥ 4-week intervals

Ferrari, et al (20)

The above are approved as an adjunct to a diet that is low in cholesterol and saturated fat.  The above agents are approved for both males and females (females must be at least one-year post-menarche) if, despite an adequate diet and other non-pharmacologic measures, the following are present: LDL-C ≥ 190 mg/dL or LDL-C ≥ 160 mg/dL and the patient has a family history of premature cardiovascular disease or two or more cardiovascular disease risk factors.
Abbreviations: mg=milligrams, LDL=low density lipoprotein

 

As Table 4 outlines, statin therapies have demonstrated variable efficacy involving clinical trials of youth (21-32). With the longest half-life, rosuvastatin is the most potent statin, followed by atorvastatin (33). Given this knowledge, if pediatric patients who are initiated on statin therapy are having trouble meeting LDL-C goals, consideration should be given to switching to rosuvastatin or atorvastatin given potency prior to exploring second-line therapy options. Simvastatin is a moderately potent statin at clinically tolerable maximum doses of 40 mg/day (33-35). Lovastatin, pravastatin, and fluvastatin, respectively, are the least potent statins (34, 35). As many studies have demonstrated, reduced potency can be compensated by an increase in the amount of statin given; however, dose escalation is often associated with an increased occurrence of adverse events (21-35). As a result, selection of a specific statin therapy should be individualized and capable of reaching treatment goals. Equally important, consideration should be given to the prevalence and severity of reported side effects (36).

 

Table 4. Statin Therapy Results

Study

Medication

Dose

Results

LDL-C

HDL-C

TC

TG

McCrindle, et al (22)

Atorvastatin

10-20 mg/day

-40%

+6%

-30%

-13%

van der Graaf, et al (23)

Fluvastatin

80 mg/day

-34%

+5%

-27%

-5%

Clauss, et al (24)

Lovastatin

40 mg/day

-27%

+3%

-22%

-23%

Lambert, et al (25)

Lovastatin

10 mg/day

-21%

+9%

-17%

-18%

Lambert, et al (25)

Lovastatin

20 mg/day

-24%

+2%

-19%

+9%

Lambert, et al (25)

Lovastatin

30 mg/day

-27%

+11%

-21%

+3%

Lambert, et al (25)

Lovastatin

40 mg/day

-36%

+3%

-29%

-9%

Stein, et al (26)

Lovastatin

10 mg/day

-17%

+4%

-13%

+4%

Stein, et al (26)

Lovastatin

20 mg/day

-24%

+4%

-19%

+8%

Stein, et al (26)

Lovastatin

40 mg/day

-27%

+5%

-21%

+6%

Knipscheer, et al (27)

Pravastatin

5 mg/day

-23%

+4%

-18%

+2%

Knipscheer, et al (27)

Pravastatin

10 mg/day

-24%

+6%

-17%

+7%

Knipscheer, et al (27)

Pravastatin

20 mg/day

-33%

+11%

-25%

+3%

Rodenburg, et al (29)

Pravastatin

20 mg/day or 40 mg/day

-29%

+3%

-23%

-2%

Wiegman, et al (28)

Pravastatin

20-40 mg/day

-24%

+6%

-19%

-17%

Avis, et al (30)

Rosuvastatin

5 mg/day

-38%

+4%

-30%

-13%

Avis, et al (30)

Rosuvastatin

10 mg/day

-45%

+10%

-34%

-15%

Avis, et al (30)

Rosuvastatin

20 mg/day

-50%

+9%

-39%

-16%

de Jongh, et al (31)

Simvastatin

10-40 mg/day

-41%

+3%

-31%

-9%

de Jongh, et al (32)

Simvastatin

40 mg/day

-40%

+5%

-30%

-17%

Adapted from National Heart Lung and Blood Institute (NHLBI): Expert Panel on Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents: Summary Report.  Pediatrics.  2011;128(S5):S213-S256: Table 9-11.
Abbreviations: mg=milligrams, LDL-C=low density lipoprotein cholesterol, HDL-C=high density lipoprotein cholesterol, TC=total cholesterol, TG=triglycerides.

 

Although long term studies evaluating the safety and efficacy of lipid-lowering medications in youth are lacking, results of short term observational and randomized controlled trials are encouraging.  For example, atorvastatin was found to be well tolerated with no statistically significant differences in adverse events reported for either the treatment or placebo groups (22). Additionally, the percentage of patients with abnormal laboratory results was similar for both groups; the only noted difference was an increased percentage of patients with elevated triglycerides in the placebo group. Treatment with atorvastatin resulted in no significant difference in sexual development as assessed by Tanner staging. Van der Graaf and colleagues found that in youth treated with fluvastatin, 58 (68.2%) reported non-serious adverse events. Only four were believed to be drug related. Treatment with fluvastatin resulted in no abnormalities in hormone levels or sexual maturation (23). 

 

In their study of lovastatin, Clauss and colleagues reported no clinically significant alterations in vital signs; growth; hormone levels including luteinizing hormone, follicle-stimulating hormone, dehydroepiandosterone sulfate, estradiol, and cortisol; menstrual cycle length; liver function tests; or muscle function tests (24). Lambert and colleagues also found lovastatin generally well tolerated with no serious clinical adverse effects noted (25). While increased, levels of aspartate aminotransferase did not exceed two times the upper limit of normal and alanine aminotransferase did not display significant changes in study participants. Creatine kinase was elevated, greater than three times the upper limit of normal, in three patients. All subjects remained asymptomatic and the elevated creatine kinase levels resolved spontaneously while no adjustment was need in their medication. Assessment of growth and sexual maturation, by Tanner staging and estimation of testicular volumes, in youth treated with lovastatin found no significant differences between the treatment groups and placebo at either 24 or 48 weeks (26). While the authors reported no significant change in serum hormone levels or biochemical parameters of nutrition, they noted that the study was under powered to detect statistically significant changes in these safety parameters.

 

Use of pravastatin has been shown to have minimal adverse events dispersed evenly between the active drug recipients and those who received placebo (27). Plasma thyroid-stimulating hormone, adrenocorticotropic hormone, cortisol, creatine phosphokinase, alanine aminotransferase, aspartate aminotransferase, total bilirubin, and alkaline phosphatase levels failed to show significant changes from baseline in all treatment groups. Rodenburg and colleagues also evaluated the safety of pravastatin based on annual or biannual evaluation of plasma creatine phosphokinase levels, liver enzymes, sex steroids, gonadotropins, and hormones of the pituitary-adrenal axis (29). Height, weight, age at menarche, Tanner staging, and testicular volume were recorded at baseline and either annually or biannually. Two subjects demonstrated elevated creatine phosphokinase levels which returned to normal without adjustments in therapy, and were presumed to be due to extreme physical exercise. No occurrence of myalgia was associated with elevation in levels of creatinine phosphokinase.  None of the subjects discontinued therapy due to adverse events or laboratory abnormalities.  Similarly, Wiegman and colleagues examined the safety of pravastatin evaluated at baseline, one year, and two years via multiple variables including: sex steroids, endocrine function parameters, height, weight, body surface area, Tanner staging, menarche or testicular volume, alanine aminotransferase, aspartate aminotransferase, and creatine phosphokinase (28). All safety parameters demonstrated no statistically significant differences between active drug recipients verses placebo in changes from baseline.

 

The safety of rosuvastatin was assessed by Avis and colleagues.  Laboratory monitoring including liver enzymes and creatine kinase levels, and markers of growth and development, such as Tanner staging (30). Two serious adverse events were reported including blurred vision in one patient in the placebo group and a vesicular rash progressing to cellulitis in one patient taking rosuvastatin 20 mg. Transaminase levels either remained normal or normalized without permanent discontinuation of treatment. While elevations in creatine kinase and reports of myalgia did occur, symptoms and creatine kinase levels normalized without permanent discontinuation of therapy. Normal progression of height, weight, and sexual development were observed. 

 

The safety of simvastatin during short term therapy has also been reported. Levels of alanine aminotransferase, aspartate aminotransferase, creatine kinase, and physical examination all demonstrated no significant differences between simvastatin treated and placebo participants (31). de Jongh and colleagues evaluated the safety of simvastatin in a second trial by monitoring adverse events as well as changes in alanine aminotransferase, aspartate aminotransferase, and creatine kinase levels (32). Of note, none of the differences reported in events or laboratory values between simvastatin recipients and placebo reached statistical significance. Additionally, there were no statistically significant differences documented for height, body mass index, cortisol levels, testicular size and testosterone levels, menstrual cycle and estradiol levels, and Tanner staging.  Dehydroepiandrosterone sulfate levels demonstrated a statistically significant decrease in the simvastatin group compared to placebo. 

 

Braamskamp and colleagues investigated the efficacy and safety of pitavastatin in pediatric patients diagnosed with hyperlipidemia. 106 patients were enrolled in the study, ages 6-17, for a 12-week period (37). Patients were randomly assigned to 4 different groups categorized by dose 1 mg, 2 mg, 4 mg, or placebo (37). The results showed a reduction in all 3 dose groups in comparison to placebo, the 1 mg group showed a 23.5% reduction in LDL-C, the 2 mg group showed a 30.1% reduction, and the 4 mg group showed a 39.3% reduction (37). There was also a 52-week extension period where patients assigned to the 1 mg group were up-titrated to a maximum dose of 4 mg to try and achieve an LDL-C level of >110 mg/dL (37). There were no safety issues of concern throughout the study. The results indicated that pitavastatin is safe and efficacious for use in pediatric patients, 6-17 years of age, and it was well-tolerated (37).

 

Long-term data regarding the impact of statin therapy on growth, development, and reduction of cardiac risk are limited, particularly for high intensity statin therapy. Recently, Kusters and colleagues evaluated the safety of statin therapy in children and adolescents with familial hypercholesterolemia after 10 years of treatment comparing laboratory safety markers as well as growth and maturation in untreated siblings (36). Only three patients discontinued therapy due to adverse events. Safety parameters such as aspartate aminotransferase, alanine aminotransferase, creatine kinase, estimated glomerular filtration rate, c-reactive protein, and age of menarche did not differ between treated patients and siblings, demonstrating safety over the 10-year treatment period. The authors do note; however, that the study was underpowered to detect the occurrence of rare events (36).

 

When initiating HMG-CoA reductase inhibitor therapy, as with any new medication therapy, it is imperative for clinicians to establish an accurate baseline, monitor for new symptoms, and counsel both patients and family members regarding potential adverse events. Females should be informed about the need to avoid pregnancy and breastfeeding while using statins. Statins may be taken with or without meals, but are commonly given with the evening meal or at bedtime as this has the potential to improve LDL-C reduction (38). As a major substrate of P450, such as CYP3A4, there are multiple drug interactions associated with statin therapy.  Grapefruit juice has gained considerable notoriety as a potential food interaction; however, it should be noted that more than a quart of grapefruit juice would have to be consumed to increase serum statin levels. Of more concern is the possible interaction of gemfibrozil and statins, which should either be avoided as it can increase the toxicity of HMG-CoA reductase inhibitors. Macrolides and antifungal azoles are classes of drugs commonly prescribed to children, and they should be avoided as often as possible as they increase serum statin levels leading to potentially enhanced myopathic effects. Additionally, patients should avoid herbal products and nutraceuticals, such as red yeast rice, which may further enhance adverse effects.

 

BILE ACID SEQUESTRANTS

 

Bile acid sequestrants, or bile acid binding resins, present an additional treatment option for youth with severe dyslipidemia. Bile acid sequestrants represent one of the oldest classes of medications available to treat dyslipidemia and were the only medication recommended in the 1992 NCEP Pediatric Panel Report, at a time when no data were available for statin use in youth (4). While no longer a recommended first-line therapy, bile acid sequestrants do have potential as a treatment option either alone or in combination with a statin (9, 10). At present, colesevelam is the only bile acid sequestrant with FDA approval for youth age 10 years and older with heterozygous familial hypercholesterolemia (39, 40). Despite the lack of FDA approval, both colestipol and cholestyramine have been studied in pediatric patients (41-49).

 

A number of clinical trials have evaluated the bile acid sequestrants in pediatric patients with heterozygous familial hypercholesterolemia and other forms of severe dyslipidemia. While palatability and tolerance remain potential barriers to effective therapy, in general, bile acid sequestrants have demonstrate significant reductions in both total cholesterol and LDL-cholesterol in study subjects (40, 43, 44, 48, 49). Table 5 provides a summary of bile acid sequestrants, pediatric approval and indications, recommended dosing ranges, comments on therapy, and supporting clinical trials. As outlined in Table 6, studies demonstrated some variability in efficacy for the available bile acid sequestrants (40, 43, 44, 48, 49).

 

Table 5. Bile Acid Sequestrants

Medication

Pediatric Approvals & Indications

Dosing

Comments

Clinical Trials

Colesevelam

Age 10-17
Heterozygous familial hypercholesterolemia

1.875 g twice daily or 3.75 g daily

May be used as monotherapy or in combination with a statin

Stein, et al (40)

Colestipol

(Note: Not FDA Approved)
Age 7-12
Primary hypercholesterolemia

5 g twice daily, 10 g daily, or
125-500 mg/kg/day

N/A

McCrindle, et al (43)
Tonstad, et al (44)

(Note: Not FDA Approved)
Age ≥12
Primary hypercholesterolemia

10-15 g/day

N/A

Cholestyramine

(Note: Not FDA Approved)
Age 6-12
Hypercholesterolemia adjunct

240 mg/kg/day divided three times daily before meals

Initiate at 2-4 g twice daily

McCrindle, et al (48)
Tonstad, et al (49)

(Note: Not FDA Approved)
Age ≥12

8 g/day divided twice daily before meals

N/A

Colesevelam is approved as an adjunct to a diet that is low in cholesterol and saturated fat.  Colesevelam is approved for both males and females (females must be at least one-year post-menarche) if, despite an adequate diet and other non-pharmacologic measures, the following are present: LDL-C ≥ 190 mg/dL or LDL-C ≥ 160 mg/dL and the patient has a family history of premature cardiovascular disease or two or more cardiovascular disease risk factors.
Abbreviations: g=grams, mg=milligrams, kg=kilograms, N/A=not applicable.

 

Table 6. Bile Acid Sequestrant Results

Study

Medication

Dose

Results

LDL-C

HDL-C

TC

TG

Stein, et al (40)

Colesevelam

1.875 g/day

-6%

+5%

-3%

+6%

Stein, et al (40)

Colesevelam

3.75 g/day

-13%

+8%

-7%

+5%

McCrindle, et al (43)

Colestipol

10 g/day

-10%

+2%

-7%

+12%

McCrindle, et al (43)

Colestipol & Pravastatin

Colestipol: 5 g/day
Pravastatin: 10 mg/day

-17%

+4%

-13%

+8%

Tonstad, et al (44)

Colestipol

2-12 g/day

-20%

-7%

-17%

-13%

McCrindle, et al (48)

Cholestyramine

8 g/day

-10% to
 -15%

+2% to +4%

-7% to
 -11%

+6% to +9%

Tonstad, et al (49)

Cholestyramine

8 g/day

-17%

+8%

-12%

N/A

Adapted from National Heart Lung and Blood Institute (NHLBI): Expert Panel on Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents: Summary Report.  Pediatrics. 2011;128(S5):S213-S256: Table 9-11.
Abbreviations: g=grams, LDL-C=low density lipoprotein cholesterol, HDL-C=high density lipoprotein cholesterol, TC=total cholesterol, TG=triglycerides.

 

Stein and colleagues assessed the safety of colesevelam at weeks 8-26 during an open-label study from.  All subjects received colesevelam 3.75 grams per day in addition to a statin (40). Safety was measured via adverse events, vital signs and physical exam, laboratory monitoring, and Tanner staging.  The most common adverse events related to use of colesevelam were gastrointestinal, including diarrhea, nausea, vomiting, and abdominal pain. It is important to note that no choking or difficulty swallowing were reported with the use of colesevelam. Vital signs, physical exams, laboratory monitoring, and Tanner staging remained the same or progressed as expected throughout the study period. 

 

McCrindle and colleagues evaluated conventional high-dose colestipol versus a combination of low-dose colestipol plus pravastatin, but did not cite safety as an endpoint for their study (40). The researchers did conduct safety monitoring in the form of laboratory tests, physical evaluation, and adverse event reporting. Significant deviations from baseline were noted for alkaline phosphatase, alanine aminotransferase, and aspartate aminotransferase at various time intervals with different medication regimens; however, the authors noted that when compared to reference values, none of the laboratory results were considered abnormal. Study participants in the two medication regimens did not significantly vary in weight gain, height changes, or body mass index. While the majority of patients experienced no adverse events, gastrointestinal symptoms such as constipation, gas or bloating, or stomach ache were more commonly reported by the patients taking the high-dose colestipol. The authors found similar suboptimal compliance with both regimens as determined by medication counts at the end of each study period. Tonstad and colleagues also assessed the tolerability of colestipol granules by monitoring side effects as well as by having subjects’ complete subjective evaluations (44). Side effects associated with colestipol included constipation, dyspepsia, flatulence, nausea, reduction in appetite, and abdominal pain. The subjective evaluations indicated that only 21% of patients liked the taste of the colestipol; however, of those who had previously taken bile acid binding resin, 86% preferred the taste of the newer orange flavored granules. Thirty seven percent of subjects also reported that they frequently forgot to take the medication, while 44% intentionally eliminated the medication from their routine on special occasions or during trips. 

 

The acceptability and compliance of cholestyramine has been studied (48). Eighty two percent (82%) of children preferred the pill formulation of cholestyramine compared to 16% who preferred the powder. Two percent of children in the study preferred neither form of the medication. Compliance was significantly impacted by medication formulation with patients taking the pill form reporting 61% compliance while those on the powder formulation were only 50% compliant. Compliance increased by at least 25% for 42% of patients when they switched to the pill formulation. Tonstad and colleagues assessed the safety of cholestyramine by measuring height velocity, erythrocyte folate, total plasma homocysteine, serum fat-soluble vitamins, and side effects (49). Weight and mean height velocity standard deviation scores were not statistically significant between treatment and placebo groups during the study. The cholestyramine active treatment group demonstrated decreased vitamin D levels and increased homocysteine levels. Differences in erythrocyte folate were not significant between the active treatment and placebo groups. Reported adverse events included intestinal obstruction, abdominal pain, nausea, and loose stools. Unpalatability was a common reason participants withdrew from the study.

 

As demonstrated by the previous studies, while bile acid sequestrants do present an effective therapy option, their side effect profile, issues of tolerability and drug interactions with statins make their use clinically challenging. It is generally recommended that all concurrent medications be given either one hour before or four hours after bile acid sequestrants to prevent decreased absorption of the additional therapies (41, 46). Use of bile acid sequestrants is generally limited to patients optimized on statin therapy who require additional therapy to achieve goal or those that cannot tolerate statins.  Data on long-term safety, however, are generally lacking. It should also be noted that bile acid sequestrants can increase triglyceride levels and should not be used in patients with increased triglyceride levels.

 

FIBRIC ACID DERIVATIVES

 

Experience with fibric acid derivatives in youth is limited. Currently there are no fibric acid derivatives with FDA approval for use in pediatric patients. Both fenofibrate and gemfibrozil are available in the United Sates, but lack pediatric data on safety, efficacy, and dosing (50, 51).  While there is very limited information on the use of bezafibrate in youth, the product is not available in the United States (52). It should be noted that fibric acid derivatives have the potential to increase the incidence in adverse events, such as rhabdomyolysis, when used with statins (45, 50, 51). However, the use of fibrates should be considered and can be beneficial in pediatric patients who also have triglyceride abnormalities (TG levels > 500 mg/dL) (9, 53).

 

NIACIN

 

Niacin provides a potential adjunct therapeutic option for youth with severe dyslipidemia who have not achieved their lipid goal. Extended-release niacin is the only formulation that has FDA-approval for use in children > 16 years of age (54). Despite a lack of FDA approval for ages younger than 16, limited efficacy and safety data are published for the use of niacin in children 10 years of age as older as adjunct therapy (54). Table 7 summarizes data on recommended dosing ranges, comments on dose adjustments, and references supporting clinical trials. 

Colletti and colleagues conducted a retrospective review to evaluate the efficacy and adverse effect profile of niacin for children with severe hypercholesterolemia (54). The effects on serum lipid profiles are detailed in Table 8.  Adverse effects were common, affecting 76% of children, and similar to those reported for adults including: flushing, abdominal pain, vomiting, headache, and elevated liver enzymes. Due to the high prevalence of adverse effects, use of niacin should be limited to patients not achieving goal with other therapies or those who cannot tolerate alternative adjunctive options. As with fibrates, niacin can also be considered for the purposes of treating pediatric patients who are concurrently diagnosed with hypertriglyceridemia (9).

 

Table 7. Niacin

Medication

Pediatric Approvals & Indications

Dosing

Comments

Clinical Trials

Niacin

Extended release: >16 years of age

Note: Immediate release is not FDA Approved:
Age ≥ 10
Adjunct therapy

Initial: 100-250 mg/day
(Max: 10 mg/kg/day)
divided three times daily with meals

May titrate weekly by 100 mg/day or every 2-3 weeks by 250 mg/day

Colletti, et al (54)

Abbreviations: mg=milligrams, kg=kilograms.

 

Table 8. Niacin

Study

Medication

Dose

Results

Colletti, et al (54)

Niacin

500-2,250 mg/day

LDL-C

HDL-C

TC

TG

-17%

+4%

-13%

+13%

Adapted from National Heart Lung and Blood Institute (NHLBI): Expert Panel on Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents: Summary Report.  Pediatrics. 2011;128(S5):S213-S256: Table 9-11.
Abbreviations: mg=milligrams, LDL-C=low density lipoprotein cholesterol, HDL-C=high density lipoprotein cholesterol, TC=total cholesterol, TG=triglycerides.

 

EZETIMIBE

 

Ezetimibe is FDA approved for adolescents 10 years of age and older with FH (53, 55), and it presents a potential therapy option either as monotherapy or when synergistically paired with an HMG-CoA reductase inhibitor (56-58). Due to its favorable tolerability, it has become the most frequently used second-line agent (59). Table 9 summarizes data on recommended dosing ranges and references supporting clinical trials while Table 10 details efficacy of therapy. 

Tolerability of ezetimibe was prospectively evaluated by Yeste and colleagues via a combination of biochemical markers and adverse event reports (56). No change was seen in hemogram, transaminases, creatinine, calcium, phosphorus, and vitamins A and E for any of the 17 patients. Additionally, there were no reports of adverse events during the study period. Clauss and colleagues retrospectively evaluated ezetimibe; therefore, safety parameters were less defined, but included intermittent measurement of liver enzymes, occasional CK levels, and adverse event reports (57). There were no reported abnormalities in liver enzymes for study participants. Ultimately, one patient was discontinued from ezetimibe therapy for asymptomatic elevated CK levels, later determined to be likely unrelated to therapy.

 

Table 9. Ezetimibe

Medication

Pediatric Approvals & Indications

Dosing

Comments

Clinical Trials

Ezetimibe

Age ≥10
Homozygous familial hypercholesterolemia

10 mg/day

N/A

Yeste, et al (56)
Clauss, et al (57)
van der Graaf, et al (58)

Abbreviations: mg=milligrams, N/A=not applicable.

 

Table 10. Ezetimibe

Study

Medication

Dose

Results

Yeste, et al (56)

Ezetimibe

10 mg/day

            LDL-C

HDL-C

TC

TG

PH

-42%

N/A

-31%

N/A

FH

-30%

-15%

-26%

N/A

Clauss, et al (57)

Ezetimibe

10 mg/day

            LDL-C

HDL-C

TC

TG

FH

-28%

N/A

-22%

N/A

FCHL

N/A

-13%

N/A

van der Graaf, et al (58)

Ezetimibe
&
Simvastatin

Ezetimibe: 10 mg/day
Simvastatin: 10-40 mg/day

            LDL-C

HDL-C

TC

TG

              -49%

+7%

-38%

-17%

                 

Adapted from National Heart Lung and Blood Institute (NHLBI): Expert Panel on Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents: Summary Report.  Pediatrics. 2011;128(S5):S213-S256: Table 9-11.
Abbreviations: mg=milligrams, LDL-C=low density lipoprotein cholesterol, HDL-C=high density lipoprotein cholesterol, TC=total cholesterol, TG=triglycerides, PH=polygenic hypercholesterolemia, FH=familial hypercholesterolemia, FCHL=familial combined hyperlipidemia.

 

Van der Graaf and colleagues assessed the safety of combination therapy with ezetimibe and simvastatin based on reported adverse events as well as laboratory monitoring and clinical examination (58). After 53 weeks, 71% of study participants reported some types of treatment-emergent adverse events. Of those events reported, only influenza, nasopharyngitis, and headache occurred in greater than 5% of participants. Consecutive transaminase elevations of at least three times the upper limit of normal were reported in 6 participants; however, all values resolved with interruption or discontinuation of therapy.  Elevations in creatine phosphokinase occurred infrequently and were not associated with myalgia at levels greater than three times the upper limit of normal.  Height, weight, and sexual maturation were not significantly impacted by therapy. Ezetimibe affords flexibility in administration time with the ability to administer it without regard to meals or time of day (45). HMG-CoA reductase inhibitors have the risk of increasing myopathy and elevation in hepatic transaminases, but are generally considered a safe combination with ezetimibe.

 

OMEGA-3 FISH OILS

 

Omega-3 fish oils are a class of therapy for which there is significantly limited data in youth. To date the FDA approved formulations of omega-3 fatty acid lack a pediatric indication. High dose omega-3 fatty acid supplementation was evaluated by de Ferranti and colleagues, but ultimately the authors found no statistically significant improvement when subjects were compared to placebo (60). Chahal and colleagues similarly found no significant impact on hypertriglyceridemia when treating pediatric patients with fish oil (61).

 

Khorshidi and associates performed a systematic review and meta-analysis of the effect of omega-3 supplementation on lipid profiles in children and adolescents. They found that omega-3 supplementation improved triglyceride levels in patients diagnosed with hypertriglyceridemia that were less than or equal to 13 years of age; however, there was no significant effect seen in HDL, LDL, or TC values (62). Omega-3 fish oils can be considered in those who have elevated triglyceride levels.

 

FAMILIAL HYPERCHOLESTEROLEMIA AND THERAPEUTIC ADVANCES

 

Familial hypercholesterolemia (FH) is a common, but often misdiagnosed inherited gene disorder (63). The most common gene mutation seen in FH is in the low-density lipoprotein receptor gene (LDLR) accounting for 85-90% of cases, followed by the apolipoprotein (ApoB) gene (5-15% of cases) and the proprotein convertase subtilisin kexin 9 (PCSK9) genes (1% of cases) (63, 64). Patients who are diagnosed with FH have abnormally elevated LDL levels from birth. FH is associated with a twenty-fold increased risk in premature cardiovascular disease and cardiovascular events (65). There are two different types of FH, heterozygous (HeFH) and homozygous (HoFH). Heterozygous patients have one mutated allele and are more commonly seen in practice, while homozygous patients have two mutated alleles and are very rare (66). Distinction between the two types is imperative because HoFH patients tend to be treatment resistant and carry a worse prognosis, if left untreated, these patients rarely make it past age 20 (66).

 

Lipid lowering in these patients can be quite challenging, especially HoFH patients. Most typical therapies for lipid lowering require functional LDL receptors, therefore given the gene mutations which often render the receptors inactive, modest reductions of 10-25% in HoFH patients are usually all that will be gained (67). HeFH patients tend to see higher rates of reduction (25-40%) (68). However, first line treatment for both forms of FH is still high-intensity statin therapy at moderate to high doses to be initiated as early as age 8 (69). All seven statins are FDA-approved for the treatment of FH and have been proven to slow down the progression of carotid intima-media thickness (63). Statins also reduce the incidence of cardiovascular events and cardiac death (63). Until recently, long term data on the use of statins in this patient population was not available, but in 2019, a 20-year follow-up study of statins in pediatric patients with FH was published (70). The results found that the incidence of cardiovascular events and death was much lower in patients treated with statins (70). If LDL-C goals are not being met with the use of statins alone, the next recommend agent is ezetimibe; however, this should not be used in patients younger than 10 years of age (59).

 

Ezetimibe is the second-line option to statins in these patients. It’s use in combination with statins has demonstrated a reduction of LDL-C levels below 135 in more than 90% of children with FH (71). In one of the only studies that assessed the coadministration of simvastatin and ezetimibe in children with HeFH, it was found to be safe, well tolerated and provided a higher LDL reduction (15%) compared with simvastatin alone in HeFH patients (72). However, when it was investigated as a monotherapy option for children with HeFH, it only produced LDL-C lowering of 27% (71, 73). It is more appropriate to use as an adjunct therapy in this population of patients.

 

Alternative medications that can be considered other second-line options are bile acid sequestrants (71). Colesevelam is safe to use, but limited to children >10 years of age (71). These drugs however have minimal LDL-C lowering effects, usually only seeing a 10-20% reduction, and more importantly are very poorly tolerated due to gastrointestinal side effects (71).

 

PCSK9 Inhibitors

 

Given the difficulty and importance of treating these patients, especially those with HoFH, there is a need for stronger lipid lowering options, which is where PCSK9 inhibitors come into play. These are a more recent class addition to the therapy options for managing FH, which help to reduce the degradation of LDL receptors and the removal of LDL-cholesterol (74). In a recent study assessing the efficacy/safety of lipid-lowering agents in patients with familial hypercholesterolemia, it was concluded that PCSK9 inhibitors were the most effective in lowering lipid levels (75). They have none of the same side effects as statins and produced similar CV benefits. Therefore, based on these conclusions, PCSK9 inhibitors are recommended as first-line agents in patients with hypercholesteremia that have intolerances or resistance to statins (75). There is currently one PCSK9 inhibitor approved for pediatric use. Repatha (evolocumab) was originally approved for use in patients 13 years of age and older (76), but the HAUSER-RCT study assessed the use of Repatha in patients ages 10-17 years of age for 24 weeks. It showed that the drug improved lipid levels (by approximately 38% in HeFH patients and 21-24% in HoFH patients) and was safe for use as the incidence of adverse events was similar in both the drug and placebo groups (71, 77). So now it is considered safe to use in pediatric patients 10 years of age and older. The HAUSER-RCT trial was then continued for another 80 weeks to further assess the safety and efficacy of Repatha (78). This new trial, HAUSER-OLE, further confirmed that the drug was safe and well-tolerated (78).

 

Praluent (Alirocumab) is another PCSK9 inhibitor that is available for treatment of FH, however it is not currently approved for use in pediatric patients (79). Nevertheless, there are studies currently assessing the safety and efficacy within this population. Bruckert and associates utilized Praluent and conducted an open-label phase 3 study specifically in pediatric patients (8-17 years of age) diagnosed with HoFH that were inadequately controlled (80). Patients received 75 or 150 mg of the drug based on weight (<50 or >50 kg, respectively) every 2 weeks for 12 weeks (80). The primary endpoint was percent change in LDL-C levels from week 0 to 12 (80). Interestingly, the results showed only a 4.1% decrease in LDL-C levels by week 12 (80). The secondary endpoints (assessing percent change LDL-C levels from baseline to weeks 24 and 48, changes in other lipid parameters from baseline to weeks 12, 24, 48, patients with a reduction of more than 15% in LDL-C levels at weeks 12, 24 and 48, and absolute change in LDL-C from baseline to weeks 12, 24, and 48) produced incredibly variable results (80). Overall, there were quite small changes in LDL-C levels observed in this study with mean reductions of LDL-C levels noted to range anywhere from ~33 to 52 mg/dL (80). More importantly, previous studies have shown that PCSK9 inhibitors are linked to a decrease in major coronary/vascular events and all-cause mortality, so although the results produced small values, the changes seen are still clinically significant based on these added benefits (80). It was also noted that this study produced similar results when compared to the ODYSSEY study which assessed the use of Praluent in adult patients with HoFH (80). The drug was deemed safe and there were no issues with tolerability (80). The study supports the use of Praluent as an adjunct therapy in HoFH patients already on first- and second-line therapies and not reaching their goal LDL-C levels (80).

 

Another study assessed the use of Praluent in pediatric patients diagnosed with HeFH. The ODYSSEY KIDS study was a phase 2 dose-finding study that enrolled pediatric patients anywhere from 8-17 years of age (81). Patients were split into 4 cohorts and dosed every 2 weeks. Dosing was determined by weight and the primary endpoint assessed percent change in LDL-C from baseline to week 8 (81). Praluent demonstrated the best reduction in LDL-C levels in the highest dosed cohorts and was well-tolerated. This study also supported the use of the drug (with further analysis) in patients who require adjunct therapy, there is a phase 3 trial planned to assess the doses from this study that resulted in the greatest reduction in LDL-C levels. Overall, it is important to note that HoFH patients are more likely to fail PCKS9 inhibitors (82). This is attributable to their mechanism of action.  This class of medication requires functional LDL receptors, and this is impaired or completely absent in HOFH patients (82). Therefore, effectiveness of PCSK9 inhibitors tends to be much higher in HeFH patients (82).

 

Leqvio (inclisiran) is also another PCSK9 inhibitor currently not approved for pediatric use (83). The mechanism of action of this drug differs from Repatha and Praluent. Leqvio is a small interfering RNA (siRNA) that utilizes the RNA interference mechanism to cause the catalytic breakdown of mRNA for PCSK9, thus stopping the translation of the protein (84). It also only requires administration twice yearly as opposed to biweekly (84). There are currently ongoing studies investigating the possibility of using Leqvio in pediatric patients. ORION-13 and ORION-16 are studies assessing the efficacy, safety and tolerability of Leqvio in pediatric patients diagnosed with HoFH and HeFH, respectively (84). They are two-part (1-year double blind, the other year open-label) phase 3 trials consisting of patients aged 12 to <18 years with FH (84). The primary endpoint is the percentage change in LDL-C from baseline to day 330 (84). Based on the results, this could be another drug option as adjunct therapy to consider for use.

 

Angiopoietin-Like Protein 3 (ANGPTL3)

 

Angiopoietin-like protein 3 (ANGPTL3) also presents a novel target of adjunctive therapy for patients with homozygous familial hypercholesterolemia that are not meeting LDL-C goals with first-line agents (85, 86).  Evkeeva (evinacumab-dgnb) is a monoclonal ANGPTL3 inhibitor that is FDA-approved specifically for the adjunctive treatment of homozygous familial hypercholesterolemia in patients 5 years of age and older (87).

 

Lomitapide

 

Lomitapide is another potential treatment option for patients with HoFH. This medication works differently from more conventional options. It binds to microsomal triglyceride transfer protein (MTP) and prevents the production of lipoproteins that contain apo-B (88). This causes a decrease in the production of very-low-density lipoprotein (VLDL) and chylomicrons. Since VLDL is converted into LDL, this mechanism ultimately causing a decrease in LDL-C levels (89). It is administered once daily at doses ranging from 5 to 60 mg (88). The side effect profile of lomitapide can be difficult for patients as it can cause severe gastrointestinal side effects (due to the decrease in absorption of fats in the intestines), most often diarrhea (89). But it is also associated with raised hepatic fats and enzymes (82). It is currently approved for adult use only, but it has become an option for use in pediatric patients through an expanded access program or a named patient basis (82). There was a case series done exploring the effect of lomitapide in 11 pediatric patients diagnosed with HoFH. It demonstrated that the drug was effective in reducing LDL-C with all 11 HoFH patients and showed a similar side effect profile to that seen in adult patients (82). GI complaints were moderated and did not cause any discontinuation of use (82). It also showed greater reduction in LDL-C levels at lower doses (82). The greatest benefit of lomitapide was associated with its ability to reduce or stop the need for lipoprotein apheresis in the patients incorporated in this case study (82). An interesting mention about lomitapide from the case series is that adult patient data shows that early intervention utilizing the drug showed a potential for increased life expectancy and a delay in the time to first major adverse cardiovascular event (82). There is also currently an ongoing phase 3, open label trial investigating the efficacy and safety of lomitapide in pediatric patients with HoFH, estimated completion date is April of 2024 (82).

 

Bempedoic Acid

 

Bempedoic acid is a new medication that exerts its effects very similarly to that of statins. It works in the same pathway as statins and targets cholesterol biosynthesis (90-92). It is administered however as a prodrug and converted to active drug only in the liver and not in the muscles (90-92). The other difference between the two classes is that bempedoic acid inhibits ATP-citrate lyase (ACL), while statins inhibit HMG CoA reductase (90-92). Due to the lack of activation in skeletal muscles, this drug is a promising alternative to patients unable to take statins due to muscle related symptoms (90-92). The medication is FDA approved for use in patients with HeFH and those with established cardiovascular disease (93). It has shown promising results in adult trials, but there are currently no published pediatric trials to date assessing the safety or efficacy of use of the drug (93). There does however appear to be a trial in development: “An Open-Label Study to Evaluate the Pharmacokinetics, Pharmacodynamics, and Safety of Bempedoic acid in Pediatric Patients with Heterozygous Familial Hypercholesterolemia.” The results are highly anticipated so that this can offer another promising drug class for use in patients intolerant or unable to meet their LDL-C goals.

 

FAMILIAL CHYLOMICRONEMIA SYNDROME

 

Familial chylomicronemia syndrome (FCS) is an incredibly rare autosomal recessive gene disorder (94). There is reduced or absent lipoprotein lipase activity causing disruption in chylomicron metabolism leading to severely elevated triglyceride levels resulting in acute recurrent pancreatitis (94). There is not however an increased risk of ASCVD with an FCS diagnosis (94). The best way to treat FCS is also often referred to as the most difficult as it requires patients to restrict dietary intake to <10-15% of daily calories (94). Other treatment options utilized are fibrates, omega-3 fatty acids and statins with variable responses, but the use of these medications is most commonly seen in patients who have multifactorial chylomicronemia syndrome (94). Given the difficulty of ensuring these patients maintain low levels of triglycerides, medications like volanesorsen are being examined (94, 95).

 

Volanesorsen

 

Volanesorsen is a second-generation 2’-O-methoxyethyl (2’-MOE) antisense therapeutic oligonucleotide. It works by inhibiting apoC3 thus lowering triglyceride plasma levels (94). When it binds to apoC3, this interrupts mRNA translation which consequently promotes triglyceride clearance/lowering of triglyceride plasma levels (94). The efficacy and safety of volanesorsen was assessed in the APPROACH study (96). It included 67 patients that were randomized to either weekly volanesorsen or placebo for 3 months (97). The results showed a 77% reduction in triglyceride plasma levels at the end of the study period and there was only 1 event of pancreatitis in the study group (97). The largest trial performed assessing the use of volanesorsen was the COMPASS trial (97). It included 114 patients who were randomized to either weekly injections of volanesorsen or placebo for a total of 26 weeks (97). The results showed that patients in the treatment group saw a reduction in triglyceride levels, chylomicron triglycerides, VLDL levels and apoC3 levels by more than 70% (97). There were also no occurrences reported of pancreatitis in any of the patients randomized to the volanesorsen group (93). In both trials, volanesorsen proved itself as a promising agent for treatment of hypertriglyceridemia in FCS patients (95-98). This drug is not approved for use in the US but is approved in other countries.

 

CONCLUSION

 

As noted in the 2011 NHLBI’s guidelines, available information regarding the treatment of youth with lipid disorders has greatly expanded. HMG-CoA reductase inhibitors, or statins, are now considered first-line pharmacologic treatment of children and adolescents with severe hypercholesterolemia who fail treatment with diet and exercise alone, although statins are only FDA approved for youth with familial hypercholesterolemia. Despite their ability to effectively reduced cholesterol levels, use of bile acid sequestrants continue to pose challenges for pediatric patients due to their unpalatability and are typically utilized as adjunctive therapy or for patients not able to tolerate statins. Fibric acid derivatives, as a class of medications, not only lack an FDA approved agent, but also continue to lack significant pediatric safety and efficacy data. Niacin, a potential adjunct therapy, lacks FDA approval for pediatric patients and is plagued by significant adverse effects, making it an unlikely therapeutic option for youth.  Ezetimibe provides clinicians with an alternative adjunct therapy option when synergistically paired with an HMG-CoA reductase inhibitor or used as monotherapy for patients intolerant to statins and bile acid sequestrants. Despite their inherit appeal and popularity amongst the lay public, omega-3 fish oils have failed to demonstrate statistically significant cholesterol lowering in pediatric and adolescent patients, but can be used to lower triglyceride levels. PCSK9 and ANGPTL3 inhibiting agents are promising novel treatment options in pediatric patients diagnosed with FH. While recent years have witnessed a dramatic increase in studies of lipid lowering medications in youth, the long-term safety and efficacy data continue to present an active focus of research.

 

REFERENCES

 

  1. McMahan CA, Gidding SS, Malcom GT, Tracy RE, Strong JP, McGill Jr HC.  Pathobiological determinants of atherosclerosis in youth risk scores are associated with early and advanced atherosclerosis. (PDAY) Pediatrics.  2006;118(4): 1447-1455.
  2. Newman III WP, Freedman DS, Voors AW, et al.  Relation of serum lipoprotein levels and systolic blood pressure to early atherosclerosis. (The Bogalusa Heart Study).  N Engl J Med. 1986;314(3):138-44.
  3. Berenson GS, Srinivasan SR, Bao W, Newman III WP, Tracy RE, Wattigney WA.  Association between multiple cardiovascular risk factors and atherosclerosis in children and young adults. (The Bogalusa Heart Study). N Engl J Med. 1998;338(23):1650-1656.
  4. National Cholesterol Education Program (NCEP): Highlights of the Report of the Expert Panel on Blood Cholesterol Levels in Children and Adolescents.  Pediatrics.  1992;89:495-501.
  5. American Academy of Pediatrics. Committee on Nutrition. Cholesterol in childhood. Pediatrics.  1998;101 (1 Pt 1): 141-7.
  6. Daniels SR, Greer FR. Lipid Screening and cardiovascular health in childhood. Pediatrics. 2008;122:198-208.
  7. National Center for Chronic Disease Prevention and Health Promotion, Division of Adolescent and School Health.  Youth Risk Behavior Sheet(YRBS): National Trends, 1991-2009: Obesity, Dietary Behaviors, and Weight Control Practices.  http://www.cdc.gov/healthyyouth/yrbs/pdf/us_obesity_trend_yrbs.pdf.  Updated October 5, 2011.  Accessed December 2011.
  8. McCrindle BW, Urbina EM, Dennison BA, Jacobson MS, Steinberger J, Rocchini AP, Hayman LL, Daniels SR. American Heart Association scientific statement: drug therapy of high-risk lipid abnormalities in children and adolescents. Circulation. 2007;115:1948-67.
  9. National Heart Lung and Blood Institute (NHLBI): Expert Panel on Integrated Guidelines for Cardiovascular Health and Risk Reduction in Children and Adolescents: Summary Report.  Pediatrics.  2011;128(S5):S213-S256.
  10. Kavey RW, Allada V, Daniels SR, Hayman LL, McCrindle BW, Newburger JW, Parekh    RS, Steinberger J. American Heart Association scientific statement: cardiovascular risk reduction in high-risk pediatric patients. Circulation. 2006;114:2710-38.
  11. Sibley C, Stone NJ. Familial hypercholesterolemia: a challenge of diagnosis and therapy. Cleve Clin J Med. 2006;73:57-64.
  12. Division of Reproductive Health, National Center for Chronic Disease Prevention and Health Promotion. US Medical Eligibility Criteria for Contraceptive Use, 2016 (US MEC). https://www.cdc.gov/reproductivehealth/contraception/mmwr/mec/summary.html. Updated March 27, 2023.Accessed November 2023.
  13. Atorvastatin product monograph. Pfizer. New York, NY. 2007.
  14. Fluvastatin product monograph. Novartis. East Hanover, NJ. 2006.
  15. Lovastatin product monograph. Merck & Co. Whitehouse Station, NJ. 2007.
  16. Pravastatin product monograph. Bristol-Myers Squibb Company. Princeton, NJ. 2007.
  17. Rosuvastatin product monograph. AstraZeneca. Wilmington, DE. 2011.
  18. Simvastatin product monograph. Merck & Co. Whitehouse Station, NJ. 2006.
  19. Pitavastatin product monograph. Kowa. Montgomery, AL. 2022.
  20. Ferrari F., Martins V.M., Rocha V.Z., Santos R.D. Advances with lipid-lowering drugs for pediatric patients with familial hypercholesterolemia. Expert Opin. Pharmacother. 2021;22:483–495. doi: 10.1080/14656566.2020.1832991.
  21. Haney EM, Huffman LH, Bougatsos C, Freeman M, Steiner RD, Nelson HD. Screening and treatment for lipid disorders in children and adolescents: systematic evidence review for the US preventive services task force. Pediatrics. 2007;120:e189-214.
  22. McCrindle BW, Ose L, Marais AD. Efficacy and safety of atorvastatin in children and adolescents with familial hypercholesterolemia or severe hyperlipidemia: a multi-center, randomized, placebo-controlled trial. J Pediatr. 2003;142:74-80.
  23. van der Graaf A, Nierman MC, Firth JC, Wolmarans KH, Marais AD, de Groot E.  Efficacy and safety of fluvastatin in children and adolescents with heterozygous familial hypercholesterolemia.  Acta Paediatr.  2006;95(11):1461-1466.
  24. Clauss SB, Holmes KW, Hopkins P, Stein E, Cho M, Tate A, Johnson-Levonas AO, Kwiterovich PO. Efficacy and safety of lovastatin therapy in adolescent girls with heterozygous familial hypercholesterolemia. Pediatrics. 2005;116(3):682-8.
  25. Lambert M, Lupien PJ, Gagné C, et al.  Treatment of familial hypercholesterolemia in children and adolescents: effect of lovastatin.  Pediatrics.  1996;97(5):619-628.
  26. Stein EA, Illingworth DR, Kwiterovich Jr PO, et al.  Efficacy and safety of lovastatin in adolescent males with heterozygous familial hypercholesterolemia.  JAMA.  1999;281(2):137-144.
  27. Knipscheer HC, Boelen CCA, Kastelein JJP, et al.  Short-term efficacy and safety of pravastatin in 72 children with familial hypercholesterolemia.  Pediatr Res.  1996;39(5):867-871.
  28. Wiegman A, Hutten BA, de Groot E, et al.Efficacy and safety of statin therapy in children with familial hypercholesterolemia: a randomized controlled trial. JAMA.  2004;292(3):331-337.
  29. Rodenburg J, Vissers MN, Wiegman A, et al.  Statin treatment in children with familial hypercholesterolemia: The younger, the better.  Circulation. 2007;116:664-668.
  30. Avis HJ, Hutten BA, Gagné C, et al.  Efficacy and safety of rosuvastatin therapy for children with familial hypercholesterolemia.  J Am Coll Cardiol.  2010;55(11):1121-1126.
  31. de Jongh S, Ose L, Szamosi T, et al.  Efficacy and safety of statin therapy in children with familial hypercholesterolemia: a randomized, double-blind, placebo-controlled trial with simvastatin. Circulation. 2002;106(17):2231-2237.
  32. de Jongh S, Lilien MR, op’t Roodt J, Stroes ES, Bakker HD, Kastelein JJ. Early statin therapy restores endothelial function in children with familial hypercholesterolemia. J Am Coll Cardiol 2002;40(12):2117-2121.
  33. Rosenson RS.  Rosuvastatin: a new inhibitor of HMG-CoA reductase for the treatment of dyslipidemia.  Expert Rev Cardiovasc Ther.  2003;1(4):495-505.
  34. Jones PH, Davidson MH, Stein EA, et al.  Comparison of the efficacy and safety of rosuvastatin versus atorvastatin, simvastatin, and pravastatin across doses (STELLAR* trial).  Am J Cardiol.  2003;92(2):152-160.
  35. Brown AS, Bakker-Arkema RG, Yellen L, et al.  Treating patients with documented atherosclerosis to National Cholesterol Education Program- recommended low-density-lipoprotein cholesterol goals with atorvastatin, fluvastatin, lovastatin, and simvastatin.  J Am Coll Cariol.   1998;32(3):665-672.
  36. Kuster DM, Avis HJ, de Groot E, et al.  Ten-year follow-up after initiation of statin therapy in children with familial hypercholesterolemia.  JAMA.  2014;312(10):1055-1057.
  37. Braamskamp MJ, Stefanutti C, Langslet G, Drogari E, Wiegman A, Hounslow N, Kastelein JJ; PASCAL Study Group. Efficacy and Safety of Pitavastatin in Children and Adolescents at High Future Cardiovascular Risk. J Pediatr. 2015 Aug;167(2):338-43.e5.doi: 10.1016/j.jpeds.2015.05.006. Epub 2015 Jun 6. PMID: 26059337.
  38. Simvastatin: pediatric drug information.  Lexicomp Online, Pediatric and Neonatal Lexi-Drugs Online, Hudson, Ohio: Lexi-Comp Inc.; 2015.
  39. Colesevelam product monograph.  Daiichi Sankyo, Inc.  Parsippany, NJ. 2009.
  40. Stein EA, Marais D, Szamosi T, et al.  Colesevelam hydrochloride: efficacy and safety in pediatric subjects with heterozygous familial hypercholesterolemia.  J Pediatr.  2010; 156(2):231-236.e1-3.
  41. Gold Standard, Inc.  Colestipol. Clinical Pharmacology [database online].  Available at: http://www.clinicalpharmacology.com. Accessed 1/4/2008.
  42. Colestipol. In: DRUGDEX® System [intranet database]. Version 5.1. Greenwood Village, Colo: Thomson Healthcare. 2007.
  43. McCrindle BW, Helden E, Cullen-Dean G, Conner WT. A randomized crossover trial of combination pharmacologic therapy in children with familial hyperlipidemia. Pediatr Res. 2002:51(6):715-21.
  44. Tonstad S, Sivertsen M, Aksnes L, Ose L.  Low dose colestipol in adolescents with familial hypercholesterolaemia.  Arch Dis Child.  1996;74(2):157-160.
  45. Taketomo CK, Hodding JH, Kraus DM, eds.  Lexi-Comp’s Pediatric Dosing Handbook.  15th edition.  Hudson, OH: Lexi-Comp; 2008.
  46. Gold Standard, Inc.  Cholestyramine.  Clinical Pharmacology [database online].  Available at: http://www.clinicalpharmacology.com.  Accessed 1/4/2008.
  47. Cholesytramine. In: DRUGDEX® System [intranet database]. Version 5.1. Greenwood Village, Colo: Thomson Healthcare. 2007.
  48. McCrindle BW, O’Neill MB, Cullen-Dean G, Helden E.  Acceptability and compliance with two forms of cholestyramine in the treatment of hypercholesterolemia in children: A randomized, crossover trial.  J Pediatr. 1997;130(2):266-273.
  49. Tonstad S, Knudtzon J, Sivertsen M, Refsum H, Ose L.  Efficacy and safety of cholestyramine therapy in peripubertal and prepubertal children with familial hypercholesterolemia.  J Pediatr.  1996;129(1):42-49.
  50. Fenofibrate. In: DRUGDEX® System [intranet database]. Version 5.1. Greenwood Village, Colo: Thomson Healthcare. 2007.
  51. Gemfibrozil.  In: DRUGDEX® System [intranet database]. Version 5.1. Greenwood Village, Colo: Thomson Healthcare. 2007.
  52. Wheeler KA, West RJ, Lloyd JK, Barley J. Double blind trial of bezafibrate in familial hypercholesterolaemia. Arch Dis Child. 1985;60(1):34-37.

 

  1. Fiorentino R, Chiarelli F. Treatment of Dyslipidaemia in Children. Biomedicines. 2021 Aug 24;9(9):1078. doi: 10.3390/biomedicines9091078. PMID: 34572264; PMCID: PMC8470054.
  2. Colletti RB, Roff NK, Neufeld EJ, Baker AL, Newburger JW, McAuliffe TL.  Niacin treatment of hypercholesterolemia in children.  Pediatrics.  1993;92(1):78-82.
  3. Ezetimibe product monograph. Merck & CO., Inc. Whitehouse Station, NJ. 2012.
  4. Yeste D, Chacon P, Clemente M, et al.  Ezetimibe as monotherapy in the treatment of hypercholesterolemia in children and adolescents.  J Pediatr Endocrinol Metab.  2009;22:487-492.
  5. Clauss S, Wai KM, Kavey REW, et al.  Ezetimibe treatment of pediatric patients with hypercholesterolemia.  J Pediatr.  2009;154:869-872.
  6. van der Graaf A, Cuffie-Jackson C, Vissers MN, et al.  Efficacy and safety of co-administration of ezetimibe and simvastatin in adolescents with heterozygous familial hypercholesterolemia.  J Am Coll Cardiol.  2008;52(17):1421-1429.
  7. Hammersley D., Signy M. Ezetimibe: An update on its clinical usefulness in specific patient groups. Ther. Adv. Chronic Dis. 2017;8:4–11. doi: 10.1177/2040622316672544.
  8. de Ferranti SD, Milliren CE, Denhoff ER, et al.  Using high-dose omega-3 fatty acid supplements to lower triglyceride levels in 10- to 19-year-olds.  Clin Pediatr (Phila).  2014;53(5):428-438.
  9. Chahal N, Manlhoit C, Wong H, McCrindle BW.Effectiveness of omega-3 polysaturated fatty acids (fish oil) supplementation for treating hypertriglyceridemia in children and adolescents.  Clin Pediatr (Phila).  2014;53(7):645-657.
  10. Khorshidi, M., Hazaveh, Z.S., Alimohammadi-kamalabadi, M. et al.Effect of omega-3 supplementation on lipid profile in children and adolescents: a systematic review and meta-analysis of randomized clinical trials. Nutr J 22, 9 (2023). https://doi.org/10.1186/s12937-022-00826-5
  11. McGowan MP, Hosseini Dehkordi SH, Moriarty PM, Duell PB. Diagnosis and Treatment of Heterozygous Familial Hypercholesterolemia. J Am Heart Assoc. 2019 Dec 17;8(24):e013225. doi: 10.1161/JAHA.119.013225. Epub 2019 Dec 16. PMID: 31838973; PMCID: PMC6951065.
  12. Diagnosis and Treatment of Heterozygous Familial Hypercholesterolemia Mary P. McGowanMD mary.p.mcgowan@hitchcock.org , Seyed Hamed Hosseini Dehkordi MD , Patrick M. Moriarty MD , and P. Barton Duell MD
  13. Defesche JC, Gidding SS, Harada-Shiba M, Hegele RA, Santos RD, Wierzbicki AS. Familial hypercholesterolaemia. Nat Rev Dis Primers. 2017 Dec 7;3:17093. doi: 10.1038/nrdp.2017.93. PMID: 29219151.
  14. Alenizi MM, Almushir S, Sulimanaa I. Surgical Management and Outcomes of Homozygous Familial Hypercholesterolemia in Two Cousins: A Rare Case Report. Cureus. 2020 Nov 25;12(11):e11692. doi: 10.7759/cureus.11692. PMID: 33391926; PMCID: PMC7769830.
  15. German, C. Homozygous Familial Hypercholesteremia: Diagnosis and Emerging Therapies. 2022 Feb 18. American College of Cardiology.
  16. Avis HJ, Vissers MN, Stein EA, Wijburg FA, Trip MD, Kastelein JJ, Hutten BA. A systematic review and meta-analysis of statin therapy in children with familial hypercholesterolemia. Arterioscler Thromb Vasc Biol. 2007 Aug;27(8):1803-10. doi: 10.1161/ATVBAHA.107.145151. Epub 2007 Jun 14. PMID: 17569881.
  17. Mirzaee S, Cameron JD. Familial hypercholesterolemia and cardiovascular risk stratification. J Clin Lipidol. 2018 Sep-Oct;12(5):1328-1329. doi: 10.1016/j.jacl.2018.07.002. Epub 2018 Jul 10. PMID: 30061010.
  18. 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 Oct 17;381(16):1547-1556. doi: 10.1056/NEJMoa1816454. PMID: 31618540.
  19. Mainieri F, Tagi VM, Chiarelli F. Recent Advances on Familial Hypercholesterolemia in Children and Adolescents. Biomedicines. 2022; 10(5):1043. https://doi.org/10.3390/biomedicines10051043
  20. van der Graaf A, Cuffie-Jackson C, Vissers MN, Trip MD, Gagné C, Shi G, Veltri E, Avis HJ, Kastelein JJ. Efficacy and safety of coadministration of ezetimibe and simvastatin in adolescents with heterozygous familial hypercholesterolemia. J Am Coll Cardiol. 2008 Oct 21;52(17):1421-9. doi: 10.1016/j.jacc.2008.09.002. PMID: 18940534
  21. Kusters DM, Caceres M, Coll M, Cuffie C, Gagné C, Jacobson MS, Kwiterovich PO, Lee R, Lowe RS, Massaad R, McCrindle BW, Musliner TA, Triscari J, Kastelein JJ. Efficacy and safety of ezetimibe monotherapy in children with heterozygous familial or nonfamilial hypercholesterolemia. J Pediatr. 2015 Jun;166(6):1377-84.e1-3. doi: 10.1016/j.jpeds.2015.02.043. Epub 2015 Apr 1. PMID: 25841542.
  22. Hajar R. PCSK 9 Inhibitors: A Short History and a New Era of Lipid-lowering Therapy. Heart Views. 2019 Apr-Jun;20(2):74-75. doi: 10.4103/HEARTVIEWS.HEARTVIEWS_59_19. PMID: 31462965; PMCID: PMC6686613.
  23. Zhao Z, Du S, Shen S, Luo P, Ding S, Wang G, Wang L. Comparative efficacy and safety of lipid-lowering agents in patients with hypercholesterolemia: A frequentist network meta-analysis. Medicine (Baltimore). 2019 Feb;98(6):e14400. doi: 10.1097/MD.0000000000014400. PMID: 30732185; PMCID: PMC6380691.
  24. Repatha product monograph. Amgen. Thousand Oaks, California. 2017.
  25. 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; HAUSER-RCT Investigators. Evolocumab in Pediatric Heterozygous Familial Hypercholesterolemia. N Engl J Med. 2020 Oct 1;383(14):1317-1327. doi: 10.1056/NEJMoa2019910. Epub 2020 Aug 29. PMID: 32865373.
  26. Santos RD, Ruzza A, Hovingh GK, Stefanutti C, Mach F, Descamps OS, Bergeron J, Wang B, Bartuli A, Buonuomo PS, Greber-Platzer S, Luirink I, Bhatia AK, Raal FJ, Kastelein JJP, Wiegman A, Gaudet D. Paediatric patients with heterozygous familial hypercholesterolaemia treated with evolocumab for 80 weeks (HAUSER-OLE): a single-arm, multicentre, open-label extension of HAUSER-RCT. Lancet Diabetes Endocrinol. 2022 Oct;10(10):732-740. doi: 10.1016/S2213-8587(22)00221-2. Epub 2022 Sep 5. PMID: 36075246.
  27. Praluent product monograph. Regeneron. Tarrytown, NY. 2021.
  28. Bruckert E, Caprio S, Wiegman A, Charng MJ, Zárate-Morales CA, Baccara-Dinet MT, Manvelian G, Ourliac A, Scemama M, Daniels SR. Efficacy and Safety of Alirocumab in Children and Adolescents With Homozygous Familial Hypercholesterolemia: Phase 3, Multinational Open-Label Study. Arterioscler Thromb Vasc Biol. 2022 Dec;42(12):1447-1457. doi: 10.1161/ATVBAHA.122.317793. Epub 2022 Nov 3. PMID: 36325897; PMCID: PMC9750107.
  29. 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 May-Jun;14(3):322-330.e5. doi: 10.1016/j.jacl.2020.03.001. Epub 2020 Mar 28. Erratum in: J Clin Lipidol. 2020 Sep - Oct;14(5):741. PMID: 32331936.
  30. Ben-Omran T, Masana L, Kolovou G, Ariceta G, Nóvoa FJ, Lund AM, Bogsrud MP, Araujo M, Hussein O, Ibarretxe D, Sanchez-Hernández RM, Santos RD. Real-World Outcomes with Lomitapide Use in Paediatric Patients with Homozygous Familial Hypercholesterolaemia. Adv Ther. 2019 Jul;36(7):1786-1811. doi: 10.1007/s12325-019-00985-8. Epub 2019 May 17. PMID: 31102204; PMCID: PMC6824397.
  31. Leqvio product monograph. Novartis. East Hanover, NJ. 2021.
  32. Reijman MD, Schweizer A, Peterson ALH, Bruckert E, Stratz C, Defesche JC, Hegele RA, Wiegman A. Rationale and design of two trials assessing the efficacy, safety, and tolerability of inclisiran in adolescents with homozygous and heterozygous familial hypercholesterolaemia. Eur J Prev Cardiol. 2022 Jul 20;29(9):1361-1368. doi: 10.1093/eurjpc/zwac025. PMID: 35175352.
  33. Kosmas CE, Bousvarou MD, Sourlas A, Papakonstantinou EJ, Genao EP, Uceta RE, Guzman E.Angiopoietin-like protein 3 (ANGPTL3) inhibitors in the management of refractory hypercholesterolemia.  Clin Pharmacol. 2022;14:49-59. doi: 10.2147/CPAA.S345072.  Epub 2022 Jul 16. PMID: 358l73366.
  34. Raal FJ, Rosenson RS, Reeskamp LF, Hovingh GK, Kastelein JJP, Rubba P, Ali S, Banerjee P, Chan K-C, Gipe DA, Khilla N, Pordy R, Weinreich DM, Yancopoulos GD, Zhang Y, Guadet D, ELIPSE HoFH Investigators. Evinacumab for Homozygous Familial Hypercholesterolemia.N Engl J Med. 2020 Aug 20;383(8):711-720. doi: 10.1056/NEJMoa2004215. PMID: 32813947.
  35. Evkeeza (Evinacumab-dgnb) product monograph. Regeneron. Tarrytown, NY. 2023.
  36. Rayan RA, Sharma S. Lomitapide. [Updated 2022 Jul 18]. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2023 Jan-. Available from: https://www.ncbi.nlm.nih.gov/books/NBK560849/
  37. Goulooze SC, Cohen AF, Rissmann R. Lomitapide. Br J Clin Pharmacol. 2015 Aug;80(2):179-81. doi: 10.1111/bcp.12612. Epub 2015 Jul 2. PMID: 25702706; PMCID: PMC4541964.
  38. Ray KK, Bays HE, Catapano AL, Lalwani ND, Bloedon LT, Sterling LR, Robinson PL, Ballantyne CM; CLEAR Harmony Trial. Safety and Efficacy of Bempedoic Acid to Reduce LDL Cholesterol. N Engl J Med. 2019 Mar 14;380(11):1022-1032. doi: 10.1056/NEJMoa1803917. PMID: 30865796.
  39. Ballantyne CM, Bays H, Catapano AL, Goldberg A, Ray KK, Saseen JJ. Role of Bempedoic Acid in Clinical Practice. Cardiovasc Drugs Ther. 2021 Aug;35(4):853-864. doi: 10.1007/s10557-021-07147-5. Epub 2021 Apr 5. Erratum in: Cardiovasc Drugs Ther. 2021 Apr 19;: PMID: 33818688; PMCID: PMC8266788.
  40. Bempedoic acid product monograph.Esperion. Ann Arbor, MI. 2022.
  41. Sunil B, Foster C, Wilson DP, Ashraf AP. Novel therapeutic targets and agents for pediatric dyslipidemia. Ther Adv Endocrinol Metab. 2021 Nov 24;12:20420188211058323. doi: 10.1177/20420188211058323. PMID: 34868544; PMCID: PMC8637781.
  42. Witztum JL, Gaudet D, Freedman SD, Alexander VJ, Digenio A, Williams KR, Yang Q, Hughes SG, Geary RS, Arca M, Stroes ESG, Bergeron J, Soran H, Civeira F, Hemphill L, Tsimikas S, Blom DJ, O'Dea L, Bruckert E. Volanesorsen and Triglyceride Levels in Familial Chylomicronemia Syndrome. N Engl J Med. 2019 Aug 8;381(6):531-542. doi: 10.1056/NEJMoa1715944. PMID: 31390500.
  43. Kolovou G, Kolovou V, Katsiki N. Volanesorsen: A New Era in the Treatment of Severe Hypertriglyceridemia. J Clin Med. 2022 Feb 13;11(4):982. doi: 10.3390/jcm11040982. PMID: 35207255; PMCID: PMC8880470.
  44. Esan O, Wierzbicki AS. Volanesorsen in the Treatment of Familial Chylomicronemia Syndrome or Hypertriglyceridaemia: Design, Development and Place in Therapy. Drug Des Devel Ther. 2020 Jul 6;14:2623-2636. doi: 10.2147/DDDT.S224771. PMID: 32753844; PMCID: PMC7351689.
  45. Gelrud A, Digenio A, Alexander V, et al. Treatment with Volanesorsen (VLN) Reduced Triglycerides and Pancreatitis in Patients with FCS and sHTG vs Placebo: results of the APPROACH and COMPASS J Clin Lipidol. 2018;12(2):537. doi: 10.1016/j.jacl.2018.03.032
  46. Kohan AB. Apolipoprotein C-III: a potent modulator of hypertriglyceridemia and cardiovascular disease. Curr Opin Endocrinol Diabetes Obes. 2015 Apr;22(2):119-25. doi: 10.1097/MED.0000000000000136. PMID: 25692924; PMCID: PMC4524519.