Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 23 August 2024

A negatively charged cluster in the disordered acidic domain of GPIHBP1 provides selectivity in the interaction with lipoprotein lipase

  • Robert Risti 1 ,
  • Mart Reimund 1 ,
  • Natjan-Naatan Seeba 1 &
  • Aivar Lõokene 1  

Scientific Reports volume  14 , Article number:  19639 ( 2024 ) Cite this article

12 Accesses

Metrics details

  • Enzyme mechanisms
  • Lipoproteins

GPIHBP1 is a membrane protein of endothelial cells that transports lipoprotein lipase (LPL), the key enzyme in plasma triglyceride metabolism, from the interstitial space to its site of action on the capillary lumen. An intrinsically disordered highly negatively charged N-terminal domain of GPIHBP1 contributes to the interaction with LPL. In this work, we investigated whether the plethora of heparin-binding proteins with positively charged regions found in human plasma affect this interaction. We also wanted to know whether the role of the N-terminal domain is purely non-specific and supportive for the interaction between LPL and full-length GPIHBP1, or whether it participates in the specific recognition mechanism. Using surface plasmon resonance, affinity chromatography, and FRET, we were unable to identify any plasma component, besides LPL, that bound the N-terminus with detectable affinity or affected its interaction with LPL. By examining different synthetic peptides, we show that the high affinity of the LPL/N-terminal domain interaction is ensured by at least ten negatively charged residues, among which at least six must sequentially arranged. We conclude that the association of LPL with the N-terminal domain of GPIHBP1 is highly specific and human plasma does not contain components that significantly affect this complex.

Similar content being viewed by others

lipoprotein lipase experiment

GIP_HUMAN[22–51] is a new proatherogenic peptide identified by native plasma peptidomics

lipoprotein lipase experiment

Structure of dimeric lipoprotein lipase reveals a pore adjacent to the active site

lipoprotein lipase experiment

Isolation of HDL by sequential flotation ultracentrifugation followed by size exclusion chromatography reveals size-based enrichment of HDL-associated proteins

Introduction.

GPIHBP1 is a capillary endothelial cell protein that transports lipoprotein lipase (LPL), the main enzyme in hydrolytic degradation of blood triglycerides, across endothelial cells to the capillary lumen. Biallelic loss-of-function mutations in GPIHBP1 cause LPL to be retained in the interstitial space, resulting in severe hypertriglyceridemia, which in turn can lead to life-threatening acute pancreatitis 1 . According to a recent study by Song et al. 2 , part of the GPIHBP1 transported LPL relocates to heparan sulfate proteoglycans (HSPGs) in the glycocalyx. At this site, LPL acts on triglyceride-rich lipoproteins, breaking down their triglycerides into fatty acids and monoglycerides that are usable components for cells. It is not known whether the displacement of LPL from GPIHBP1 to the glycocalyx is entirely spontaneous or is influenced by some plasma component. Understanding this mechanism may be important because its misregulation can lead to elevations in plasma triglyceride levels, which are correlated with an increased risk of cardiovascular disease and other metabolic disorders.

Two distinct regions of GPIHBP1—an intrinsically disordered N-terminal acidic domain with a long stretch of aspartate and glutamate residues, and a cysteine-rich Ly6 domain with a three-finger structure – are involved in the interaction with LPL. Both GPIHBP1 domains are also required for its transport function 3 . Surface plasmon resonance (SPR) studies show that while the interaction of the N-terminal domain with LPL is characterized by a high on–off rate, the complex with Ly6 forms and also dissociates at a slower rate 4 , 5 , 6 . While the sequence of the Ly6 domain is highly conserved between species, the N-terminal domains differ in length, distribution, and number of negative charges 7 . For example, human N-terminal domain of GPIHBP1 contains 21 negatively charged residues, but mouse or opossum have 17 or 32 acidic residues, respectively. X-ray structures of the LPL/GPIHBP1 complex reveal that the Ly6 domain binds to the C-terminal domain of LPL and that this association is mainly hydrophobic in nature 8 , 9 . Although the binding region of LPL involved in the interaction with the N-terminal domain was not identified in the X-ray structures, it is probable that the highly positively charged region of LPL plays a crucial role in its formation. This region is also likely to be involved in LPL’s interaction with other polyanions such as the polysaccharides heparin and heparan sulfates 10 . Comparison of affinities of synthetic peptides corresponding to the sequence of the N-terminal domain of mouse, bovine or human GPIHBP1 suggested that the number of negatively charged residues in the N-terminal domain might be important 4 . The GPIHBP1 N-terminal domain of several species also contains a tyrosine whose sulfation may play a role in affinity for LPL 6 . However, the structural motifs of the N-terminal domain of GPIHBP1 that confer its high affinity to LPL remain largely unclear.

Previous studies have shown that LPL binds strongly to heparin and heparan sulfates and, like the N-terminal domain, this interaction is transient and characterized by high on- and off-rates 11 . Furthermore, the interaction seems to not be highly specific, LPL prefers HS sequences with high negative charge density 12 , 13 . These observations raise a question of specificity for the binding of the N-terminal domain to LPL. Is it completely non-specific, having only an additional role in the interaction with full-length GPIHBP1 or does it play a specific role? It is noteworthy to bear in mind that many heparin-binding proteins have been identified in human plasma 14 , 15 , 16 , the common feature of which is the presence of a positively charged region in their folded structure 17 . Considering that the interaction of some proteins with heparin is primarily non-specific and ionic, depending mainly on the charge density, it is reasonable to ask whether the N-terminal domain of GPIHBP1 could be a potential ligand for them.

Here, we investigated whether human plasma contains components that bind to the N-terminal domain of GPIHBP1 or affect its interaction with LPL. Using SPR, affinity chromatography, and fluorescence techniques, we were unable to identify any plasma component, other than LPL, that bound the N-terminus with detectable affinity or affected its interaction with LPL. In contrast, we ascertain several plasma proteins that bound avidly to heparan sulfate or heparin. Based on these observations, we conclude that the association of LPL with the N-terminal domain of GPIHBP1 is highly specific and human plasma does not contain components that substantially affect this complex. By combining binding and stabilization studies of different synthetic peptides, we show that the high affinity of the LPL/N-terminal domain interaction is ensured by at least ten negatively charged residues, among which at least six must be in a sequentially arranged cluster.

Materials and methods

LPL was purified from bovine milk as previously described 18 . Aliquoted stock solutions of LPL were stored at −80 °C in a buffer containing 10 mMBis-Tris, pH 6.5, 1.5 M NaCl. Samples were used immediately after thawing and only used once. Synthetic peptides (Table 1 ) were purchased from GeneCust (Luxembourg). Some peptides contained an extra cysteine residue at the C-terminal end. The synthetic peptides were biotinylated and labeled by DyLight 488 dye (Pierce) at cysteines as previously described 4 . Heparan sulfate was biotinylated at amino groups as previously described 11 . Extinction coefficients at 280 nm for determination of peptide/protein concentrations were as follows: LPL—70,440 M −1 cm −1 ; human N-terminal peptide, bovine N-terminal peptide, peptide 1 and peptide 4–1480 M −1 cm −1 ; mouse N-terminal peptide—2960 M −1 cm −1 . The extinction coefficients were calculated according to Gill and Hippel 19 . Concentrations of LPL were calculated using its monomer molecular weight of 55 kDa. Heparin binding proteins, namely antithrombin III (#194,936) and fibroblast growth factor 2 (FGF-2) (#153,509) were purchased from ICN Biomedicals Inc., protamine (#P4005) was purchased from Sigma Aldrich. 1,2-Di-O-lauryl-rac-glycero3-glutaric acid 6-methylresorufin ester (DGGR) (#30,058) was purchased from Sigma Aldrich. Nonfasting human EDTA-plasmas were purchased from the Tallinn Blood Centrum. Lipoprotein deficient human plasma was obtained from plasma by using an ultracentrifuge as previously described 20 . Triglyceride and cholesterol concentrations were determined using Triglyceride Colorimetric Assay Kit (Cayman, USA) and Cholesterol Fluorometric Assay Kit (Cayman, USA).

Surface plasmon resonance experiments

SPR experiments were conducted on a Biacore 3000 (GE Healthcare Life Sciences) instrument. Neutravidin (Sigma) was covalently attached to the surface of CM5 (GE Healthcare Life Sciences) sensor chips using the amine coupling kit (GE Healthcare Life Sciences). To study the binding of human plasma components to different ligands, biotinylated N-terminal peptide of GPIHBP1 or biotinylated heparan sulfate were bound to the sensor chip surface via biotin-NeutrAvidin interaction. 578 RU of biotinylated N-terminal GPIHBP1 was bound to the surface of the sensor which corresponded to a surface density of 148.5 fmol/mm 2 .

Nonfasting pooled human plasma samples with a mean TG concentration of 1.08 mM ( n  = 8) or pooled lipoprotein deficient human plasma samples ( n  = 3) were diluted tenfold and injected over the surfaces (60 µl, 20 µl/min). In control experiments, 20 nM, 40 nM or 80 nM purified bovine LPL was added to the same tenfold diluted pooled lipoprotein deficient human plasma. Measurements were performed at 25 °C in a running buffer containing 20 mM HEPES, pH 7.4, 0.15 M NaCl. The same measurement conditions were also used in the control experiments with isolated heparin binding proteins (antithrombin III, protamine, FGF-2).

The affinity between various peptides and LPL was assessed using a SPR competition assay, following the experiments first described by Reimund et al. 4 . The sequences of the investigated peptides are presented in Table 1 . The experimental steps of the binding study were as follows: (1) Biotinylated N-terminal peptide of GPIHBP1 was immobilized to the sensor chip via NeutrAvidin. (2) LPL (800 nM) was mixed with different peptides at increasing concentrations at 4 °C. LPL or solutions of LPL/peptide complexes were injected over the sensor chip’s surface (30 µl, 5 µl/min). (3) Binding of LPL to the immobilized N-terminal peptide of GPIHBP1 was registered near the equilibrium at the end of each injection. Measurements were carried out at 4 °C. Running buffer contained 20 mM NaH 2 PO 4 , pH 7.4, 2 mg/ml BSA, 0.4 M NaCl. The affinity of the peptides for LPL, expressed as K D values, was calculated using Eq. ( 1 ).

where P 0 is the concentration of peptide in the injected solution, the K D is the equilibrium dissociation constant, Δ R is the change in SPR at equilibrium, and ‘ a ’ is the change in SPR for binding of LPL to the surface when the solution did not contain free peptide.

Direct binding studies were performed to investigate the interaction between peptide 6 and LPL (Table 1 ). Biotinylated peptide 6 was attached to the surface via NeutrAvidin and solutions of 1, 2 or 3 µM LPL were injected over this surface. Experimental conditions were the same as in competition experiments.

Affinity chromatography

Fasting human plasma (17 ml) was loaded to an 8 ml heparin column (Heparin Sepharose 6 Fast Flow affinity resin, GE Healthcare Life Sciences) in a buffer containing 100 mM TRIS, pH 7.4, 0.15 M NaCl. The column was washed with 80 ml of the same buffer. Heparin binding proteins (i.e. bound material) were eluted from the column using a buffer containing 100 mM TRIS, pH 7.4, 2 M NaCl. Eluted proteins were pooled together, dialyzed to buffer containing 100 mM TRIS, pH 7.4, 0.15 M NaCl and loaded to an affinity column containing immobilized N-terminal peptide of GPIHBP1. The latter was made by attaching biotinylated N-terminal peptide to HiTrap Streptavidin HP column (GE Healthcare Life Sciences). The column was washed with 100 mM TRIS, pH 7.4, 0.15 M NaCl buffer and elution was performed in a 0.15–2 M NaCl gradient. All steps were carried out at 10 °C. Identification of human plasma proteins that bound to the heparin column was obtained as a service from the Proteomics Core Facility at the Institute of Technology, University of Tartu (Tartu, Estonia).

Fluorescence anisotropy

Fluorescence anisotropy experiments were conducted on a Hitachi F-7000 (Hitachi High-Tech, Japan) fluorescence spectrophotometer. The excitation and emission wavelengths were 493 nm and 518 nm, respectively. Experiments were done either in phosphate buffer (20 mM phosphate, pH 7.4, 0.15 M NaCl), or in the same buffer with added 10 mg/ml BSA, 10 IU/ml heparin or 50% lipoprotein free human plasma. Dylight 488 maleimide labeled N-terminal peptide of GPIHBP1 at concentrations of 100 nM were mixed with increasing concentrations of LPL. Next, fluorescence was measured when excitation and emission polarizers both were oriented vertically, and when the excitation polarizer was oriented vertically, and the emission polarizer was oriented horizontally.

From the acquired data the fluorescence anisotropy was calculated using Eq. ( 2 ):

where r is fluorescence anisotropy, \({I}_{\parallel }\) is observed fluorescence intensity when the emission polarizer was oriented parallel to the direction of the polarized excitation, \({I}_{\perp }\) is the observed fluorescence intensity when the emission polarizer was oriented perpendicular to the direction of the polarized excitation.

Dependence between change of anisotropy and LPL concentration was fitted with Eq. ( 3 ) for calculation of the K D values for a 1:1 binding model:

where r is the change of fluorescence anisotropy, r max is the value of anisotropy of the complex of LPL with the N-terminal peptide, L 0 is the concentration of LPL, P 0 is the concentration of the peptide, K D is the equilibrium dissociation constant.

Determination of LPL activity with DGGR

Fluorescence spectroscopy was utilized to assess the effects of the N-terminal GPIHBP1 and synthetic peptides on the thermostability of LPL by monitoring LPL activity using the fluorogenic substrate DGGR. Experiments were conducted using a spectrofluorophotometer (Shimadzu RF-5301 PC, Shimadzu Corporation, Japan) at an excitation wavelength of 572 nm and an emission wavelength of 605 nm. 10 nM LPL was incubated at 37 °C in 20 mM HEPES, pH 7.4, 150 mM NaCl either alone, with 1 μM nGPIHBP1, or with 1 μM synthetic peptides. Initial LPL activity was measured at 37 °C immediately after mixing, and then at 10-min intervals.

Modeling of the predicted complex

The cryo-EM structure of LPL for the visualization of positively charged surfaces was obtained from the PDB entry 8ERL 21 . The structure of LPL in complex with either full-length GPIHBP1 or the N-terminal domain of GPIHBP1 was predicted using ColabFold 22 , which uses an optimized approach to accelerate protein–protein complex modeling by Alphafold2-Multimer 23 . The human LPL and GPIHBP1 sequences were obtained from Uniprot database entries Q6IAV0 and Q8IV16, respectively 24 . Five predicted structures were generated for each complex based on the chosen sequences and the model with the highest confidence score was chosen for visualization. ChimeraX was used to create the figures and calculate the Coulombic electrostatic potential of predicted structure surfaces 25 .

LPL is the only plasma component that binds the N-terminal domain of GPIHBP1 with high affinity

To investigate whether human plasma contains proteins in addition to LPL that bind to the N-terminal domain of GPIHBP1, we performed binding studies using SPR. Solutions of pooled human plasma samples ( n  = 8) or pooled lipoprotein-deficient human plasma samples ( n  = 3) were injected into sensor chip flow cells pre-immobilized with the N-terminal domain of GPIHBP1 (Fig.  1 ) or with heparan sulfate (Fig.  2 ). To assess the effect of nonspecific binding, the plasma solutions were injected into the flow cells that did not contain the N-terminal domain or heparan sulfate. The results of these experiments suggested that neither lipoprotein-deficient nor lipoprotein-containing human plasma contained components that bound to the N-terminal domain of GPIHBP1 with detectable affinity. In the case of lipoprotein-containing plasma, nonspecific binding to the sensor chip surface was even higher than to the N-terminal domain (Fig.  1 b). However, when 20 nM, 40 nM or 80 nM LPL was present in the lipoprotein deficient plasma samples, association with the N-terminal domain of GPIHBP1 was evident, and increased when the concentration of LPL was raised (Fig.  1 c–e). At the same time, injecting various heparin-binding proteins (antithrombin III, protamine, FGF-2) over the N-terminal domain of GPIHBP1 showed no binding, even at concentrations of 1 μM (Supplementary Fig. S1 ).

figure 1

SPR studies assessing the binding of human plasma components to the N-terminal peptide of GPIHBP1. Solutions of tenfold diluted lipoprotein free human plasma ( a ), tenfold diluted whole plasma ( b ), or the same lipoprotein free plasma with added 20 nM LPL ( c ), 40 nM LPL ( d ) or 80 nM LPL ( e ) were injected to the flow cells with the immobilized N-terminal peptide of GPIHBP1. Solid line shows specific binding to the N-terminal peptide of GPIHBP1. Dashed line shows nonspecific binding to sensor chip matrix.

figure 2

SPR studies assessing the binding of human plasma components to heparan sulfate. Solutions of tenfold diluted lipoprotein free human plasma ( a ), tenfold diluted whole plasma ( b ), or the same lipoprotein free plasma with added 20 nM LPL ( c ) were injected to the flow cells with the immobilized heparan sulfate. Solid line shows specific binding to heparan sulfate. Dashed line shows nonspecific binding to sensor chip matrix.

In contrast to the N-terminus of GPIHBP1, considerable binding of human plasma components to the immobilized heparan sulfate was observed (Fig.  2 a,b), consistent with previous studies showing that human plasma contains numerous heparin-binding proteins 14 , 15 , 16 . These experiments suggested that even though human plasma contains many heparin binding proteins, they either do not interact with the N-terminal domain of GPIHBP1, or their affinity to this domain is much lower than that of LPL.

In addition to SPR experiments, we used affinity chromatography to study the binding of plasma components to the N-terminal domain of GPIHPB1 and heparin. We used a streptavidin-Sepharose column onto which biotinylated N-terminal of GPIHBP1 was immobilized. Heparin was directly immobilized to CNBr activated agarose. Mass spectrometry identified a total of 76 plasma proteins, which were bound to heparin-agarose and subsequently eluted using a NaCl concentration gradient. The strongest proteolytic peptide intensity signals were given by antithrombin, kallikrein, thrombin, apolipoprotein E, histidine rich glycoprotein, alpha1 microglobulin, coagulation factor XI, complement factor D. However, these proteins did not bind to the N-terminal domain of the GPIHBP1 affinity column (Fig.  3 a)—all these proteins were present in the flow-through and no additional proteins were eluted in the NaCl gradient. The same was observed when whole plasma was injected over the affinity column (Fig.  3 b). In contrast, LPL in the presence of BSA bound to the column with immobilized N-terminal GPIHPB1 and eluted at NaCl concentration between 0.4 and 0.6 M (Fig.  3 c).

figure 3

Testing the ability of heparin binding proteins purified from human plasma to bind to the N-terminal domain of GPIHBP1 using affinity chromatography. Biotinylated N-terminal peptide was attached to HiTrap Streptavidin HP column. ( a ) Plasma proteins eluted from heparin affinity column or ( b ) whole human plasma was loaded to the column. All proteins were in the flow-through. No additional proteins eluted in 0.15–2 M NaCl gradient. ( c ) LPL was loaded to the N-terminal peptide column and eluted in a 0.15–2 M NaCl gradient. Solid line shows optical density at 280 nm. Dashed line shows NaCl concentration.

Components of human plasma do not substantially interfere with the interaction between LPL and the N-terminal domain of GPIHBP1

To further investigate the interaction between LPL and the N-terminal domain of GPIHBP1, we used fluorescence anisotropy which allowed to conduct studies in undiluted blood plasma, i.e. the environment in which LPL functions in vivo. For comparison, measurements were performed under four different conditions: (1) in a standard phosphate buffer, (2) in the same buffer with added 10 mg/ml BSA, (3) in the same buffer with 10 IU/ml heparin, or (4) in the same buffer which contained lipoprotein free human plasma. As can be seen in Fig.  4 , the affinity of the interaction was the highest in phosphate buffer ( K D  = 60 nM). In buffers with BSA or lipoprotein free plasma the interaction was slightly weaker. Introducing heparin to the buffer completely blocked the interaction as heparin is known to disrupt the N-terminal GPIHBP1-LPL complex 4 . However, in these cases a simple binding model did not fit the data well and the K D value could not be reliably determined. This suggests a more complex binding mechanism. The comparable effect of BSA and plasma on the interaction suggests that albumin acted as an effector, whereas other components of human plasma did not appear to influence this interaction. The weak binding of BSA to LPL as shown in our recently published study would explain the slightly reduced affinity and the complex interaction mechanism 26 . These observations are in line with SPR findings in diluted human plasma (Fig.  1 ) showing the specificity and selectivity of this interaction and suggest that other plasma components do not compete with LPL to interact with the N-terminal domain of GPIHBP1.

figure 4

Human plasma environment does not influence the affinity of the interaction between the N-terminal domain of GPIHBP1 and LPL. The change of fluorescence anisotropy of labeled N-terminal domain of GPIHBP1 was measured at different concentrations of LPL in standard phosphate buffer (black), in the buffer with 10 mg/ml BSA (white), or with 10 IU/ml heparin (blue), or in the buffer with 50% lipoprotein free human plasma (red).

A cluster of negatively charged residues drives specificity of the interaction between the N-terminal domain of GPIHBP1 and LPL

The observations that binding of LPL to the N-terminal domain of GPIHBP1 is specific prompted us to ask what structural features of the N-terminal domain ensure the high affinity of this interaction (Fig.  5 ). Firstly, we examined whether various regions of the domain may have different contributions to the total affinity. We chose peptides 1 and 2 (see Table 1 ), whose sequences partially overlapped but represented distinct regions within the human N-terminal domain of GPIHBP1. These peptides contained the same number (10) of negatively charged residues. Their affinities for LPL were compared to peptide 3 which also contained 10 negatively charged residues and whose affinity has been already previously determined 4 . Together the sequences of these three peptides make up the entire N-terminal domain sequence. All three peptides bound to LPL with affinities that did not differ greatly. However, peptide 3 had the highest affinity and peptide 1 had the lowest affinity. Their comparable affinities for LPL suggested that all parts of the N-terminal domain of GPIHBP1 contribute to the interaction with LPL.

figure 5

The affinities for the interaction between LPL and different peptides as determined by SPR. Peptides 1–5 sequences correspond to the various regions of the human N-terminal domain of GPIHBP1. Peptides 6 and 7 sequences are partially based on the human N-terminal domain of GPIHBP1. Errors shown are S.D. of the fitting. K D values for human N-terminal peptide, peptide 3, peptide 4, peptide 5, peptide 7, mouse N-terminal peptide and bovine N-terminal peptide have been published previously 4 .

Next, we examined how important is the number and distribution of negatively charged groups. We designed peptide 6, the sequence of which was obtained by replacing several aspartate and glutamate residues with alanine in the sequence of the human N-terminal domain of GPIHBP1. As a result of such substitutions, there were no regions in peptide 6 with more than two adjacent negatively charged residues. The 21 amino acid sequence of peptide 6 contained 13 negatively charged groups, as did peptide 5, whose negatively charged residues were arranged compactly and formed a cluster of seven negatively charged residues in the middle of its 15 amino acid sequence. Peptide 5 exhibited a binding affinity to LPL comparable to that of the full-length N-terminus. Conversely, peptide 6 failed to demonstrate detectable binding under the experimental conditions employed. These observations suggest either considerably reduced affinity to LPL or a complete lack of interaction. Thus, the presence of only a certain number of negatively charged residues alone does not guarantee high affinity binding to LPL. Instead, the sequential arrangement of the negatively charged group in the sequence emerges as the critical factor. At the same time, the presence of negatively charged clusters was also not sufficient, because short cluster-containing peptides 4 and 8 did not interact with LPL. Summarizing the obtained results on Fig.  5 , our findings indicate that for robust binding to LPL, it is essential for a sequence to contain at least one negatively charged cluster of 6 residues, along with a minimum total of 10 negatively charged residues.

Clusters of negatively charged residues are crucial for LPL stabilization

As previous studies have shown, the N-terminal domain of GPIHBP1 stabilizes LPL from spontaneous thermal inactivation 5 , 27 . We therefore investigated whether the same peptides that were used in the interactions study could also prevent loss of catalytic activity of LPL at 37 °C (Fig.  6 and Supplementary Fig. S2 ). Peptide 4, which has the shortest sequence and no detectable affinity to LPL, could not stabilize LPL compared to plain buffer. Peptide 8, with two additional negatively charged residues compared to peptide 4, also failed to exert any effect on the stability of LPL. However, as the number of negatively charged residues increased further, so did the effect of the peptides on the stability of LPL. Peptide 3 conserved 60% of LPL activity. Peptide 5, and the human N-terminal peptide, which had a higher affinity to LPL compared to peptide 3, increased LPL stability even further. Interestingly, while peptide 6 has the same number of negatively charged residues as peptide 5, it failed to increase LPL stability compared to plain buffer. This demonstrates that, much like the binding affinity of peptides to LPL is not solely dependent on the number of negatively charged residues, neither is the effect of peptides on LPL stability.

figure 6

Clusters of negatively charged residues are crucial for LPL stabilization. 10 nM LPL was incubated at 37°C with 1 μM nGPIHBP1 or peptides. LPL activity was determined with DGGR immediately, and after 10 min of incubation. Results are presented as a mean percentage of remaining initial activity ± SD of three independent measurements. LPL stability was highly dependent on the presence of negatively charged amino acid clusters, and their size.

Predicted models show interaction of the N-terminal domain of GPIHBP1 with the large positively charged region of LPL

LPL has a large positively charged region spanning from the C-terminal domain, across the hinge region, to the N-terminal domain as shown on Fig.  7 a. This patch is comprised of four distinct positive charge clusters of which three are important for binding to heparin 10 . We hypothesized that the same clusters might be important for the interaction with the N-terminal domain of GPIHBP1. While the binding of the Ly6 domain of GPIHBP1 has been elucidated 8 , 9 , the binding of the N-terminal domain has remained elusive, likely due to its disordered and dynamic nature. We therefore used ColabFold to predict the localization of the interaction interface between the N-terminal domain of GPIHBP1 and LPL (Fig.  7 b). All predicted models consistently demonstrated binding of the highly negatively charged N-terminal domain of GPIHBP1 to the basic patch of LPL. This was also true in the case of full-length GPIHBP1 (Fig.  7 c), where the Ly6 domain was simultaneously bound to the same C-terminal region of LPL as shown in crystal structures 8 , 9 . In both complexes shown in Fig.  7 , the C-terminal segment of the N-terminal domain, which contains a cluster of negatively charged groups (DEEDEDEVEEEE), directly interacts with the surface of LPL. The rest of the domain remains away and can support interactions with long-range electrostatic forces. However, it should be considered that the per-residue model confidence score (pLDDT) for the predicted structure of the complexes was very low (< 50), which means that interactions between other regions are also possible. It is also feasible that the low scores are a consequence of the inherent disorder of the N-terminal GPIHBP1 peptide and are a predictor of the dynamic nature of the peptide, rather than a sign of low prediction confidence 28 .

figure 7

Predicted models showing how the N-terminal domain of GPIHBP1 interacts with the positively charged patch of LPL. ( a ) LPL has a wide polyanion binding region as visualized on an experimentally derived structure. Predicted complexes of LPL with the N-terminal domain of GPIHBP1 ( b ) or with full-length GPIHBP1 ( c ) demonstrate how the highly negatively charged region (red) of GPIHBP1 fits into the large positively charged patch (blue) of LPL. The N-terminus of GPIHBP1 is indicated.

In the present study, we show that although plasma contains numerous proteins that bind to the negatively charged polysaccharides heparin and heparan sulfate, we were unable to identify any that interacted with the highly negatively charged and disordered N-terminal domain of GPIHBP1 with detectable affinity. At the same time LPL binds to this domain with high affinity 4 , 6 . Furthermore, plasma proteins did not block the interaction between LPL and the N-terminal domain in tenfold diluted plasma as shown by SPR. This observation supports the speculation that movement of LPL from GPIHBP1 to HSPGs in the glycocalyx might occur spontaneously 2 . Fluorescence anisotropy measurements in undiluted human plasma suggested that only albumin, the major plasma protein, slightly reduced this interaction. Such a large difference between LPL and other heparin-binding proteins was somewhat unexpected because several of the plasma proteins interact with heparin with considerable affinity 17 . Moreover, the interaction of several of these proteins with heparin has been shown to be rather non-specific without favoring specific heparin sequences. In addition, plasma concentrations of some heparin-binding plasma proteins are relatively high. A prime illustration of this concept can be found in thrombin, which typically maintains a plasma concentration ranging between 5 and 10 mg/dl or 0.7–1.3 µM 29 and its interaction with heparin is predominantly nonspecific 30 . Thus, it is difficult to find an explanation for the large difference between LPL and other heparin-binding proteins. We speculate that the high affinity of LPL/N-terminal domain interaction is due to the large positively charged polyanion binding region in LPL ( ∼ 2400 Å 2 ) 8 , which includes the region in the C-terminal domain, the hinge region, and the N-terminal catalytic domain of the enzyme. In other heparin-binding proteins, the area of the corresponding regions is substantially smaller, ranging from 200 to 1150 Å 17 . The wide ( ∼ 60 × 40Å 2 ) 8 and positively charged region in LPL may allow the N-terminal domain to be positioned there in several different ways, leading to more degrees of freedom, an increase in binding entropy and thus a potential increase in the binding affinity 31 , 32 .

The results of the present study suggest that the high affinity of the N-terminal domain for LPL is ensured by the presence of at least 10 negatively charged residues, at least 6 of which are in a consecutive cluster. This conclusion is also supported by the knowledge that all known GPIHBP1 sequences from different species have both a cluster and a sufficient number of negative residues in their N-terminal domain. The important role of the cluster was particularly evident in the comparison of peptides 5 and 6 which had an equal number of negative charges but showed a considerable difference in their interaction with LPL. Only the cluster-containing peptide 5 bound to LPL with substantial affinity and stabilized its active conformation. According to the small-angle X-ray scattering (SAXS) analyses 6 , the chain length of the disordered N-terminal domain is 68 Å, most of which could fit into the highly positively charged polyanion-binding region in LPL. However, our results suggest that the high density of negative charge of the cluster region is a prerequisite for strong binding. Therefore, the cluster region is likely to be the "hot" region of the N-terminal domain in the interaction with LPL. Our data further indicates a strong correlation between the affinity of the examined peptides and their capacity to stabilize LPL. It is reasonable to infer that the stabilization of LPL necessitates a dense concentration of negatively charged sequences within the polypeptide chain, facilitating binding to regions of LPL characterized by a high density of positive charge. Electrostatic repulsion forces likely underlie LPL's instability, arising from positively charged residues within LPL's highly positively charged region.

Our modeling findings align with the SPR experiments and stabilization studies: the cluster situated at the C-terminus of the N-terminal domain EEDEDEVEEEETC is projected to establish direct contact with the LPL region positioned at the junction of the C-terminal domain and hinge region. Other segments within the N-terminal domain are not directly engaged. Their influence on the interaction likely stems from long-range electrostatic effects. All examined peptides demonstrated interaction with this region in the modeled complexes.

In summary, in this report we show that the interaction between the strongly negatively charged disordered N-terminal domain of GPIHBP1 and the natively folded LPL is highly specific in human plasma environment. The sequential cluster of negatively charged residues ensures this specificity. Other plasma components do not substantially influence this interaction.

Data availability

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Davies, B. S. et al. GPIHBP1 is responsible for the entry of lipoprotein lipase into capillaries. Cell Metab. https://doi.org/10.1016/j.cmet.2010.04.016 (2010).

Article   PubMed   PubMed Central   Google Scholar  

Song, W. et al. The lipoprotein lipase that is shuttled into capillaries by GPIHBP1 enters the glycocalyx where it mediates lipoprotein processing. Proc. Natl. Acad. Sci. https://doi.org/10.1073/pnas.2313825120 (2023).

Song, W. et al. Electrostatic sheathing of lipoprotein lipase is essential for its movement across capillary endothelial cells. J. Clin. Investig. https://doi.org/10.1172/jci157500 (2022).

Reimund, M. et al. Evidence for two distinct binding sites for lipoprotein lipase on glycosylphosphatidylinositol-anchored high density lipoprotein-binding protein 1 (GPIHBP1). J. Biol. Chem. 290 (22), 13919–13934. https://doi.org/10.1074/jbc.m114.634626 (2015).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Mysling, S. et al. The acidic domain of the endothelial membrane protein GPIHBP1 stabilizes lipoprotein lipase activity by preventing unfolding of its catalytic domain. eLife. https://doi.org/10.7554/eLife.12095 (2016).

Kristensen, K. et al. A disordered acidic domain in GPIHBP1 harboring a sulfated tyrosine regulates lipoprotein lipase. Proc. Natl. Acad. Sci. U.S.A. https://doi.org/10.1073/pnas.1806774115 (2018).

Holmes, R. & Cox, L. Comparative studies of glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1: evidence for a eutherian mammalian origin for the GPIHBP1 gene from an LY6-like gene. 3 Biotech. https://doi.org/10.1007/s13205-011-0026-4 (2012).

Birrane, G. et al. Structure of the lipoprotein lipase–GPIHBP1 complex that mediates plasma triglyceride hydrolysis. Proc. Natl. Acad. Sci. U.S.A. https://doi.org/10.1073/pnas.1817984116 (2019).

Arora, R. et al. Structure of lipoprotein lipase in complex with GPIHBP1. Proc. Natl. Acad. Sci. U.S.A. https://doi.org/10.1073/pnas.1820171116 (2019).

van Tilbeurgh, H., Roussel, A., Lalouel, J. & Cambillau, C. Lipoprotein lipase. Molecular model based on the pancreatic lipase x-ray structure: consequences for heparin binding and catalysis. J. Biol. Chem. https://doi.org/10.1016/S0021-9258(17)41822-9 (1994).

Article   PubMed   Google Scholar  

Lookene, A., Chevreuil, O., Østergaard, P. & Olivecrona, G. Interaction of lipoprotein lipase with heparin fragments and with heparan sulfate: Stoichiometry, stabilization, and kinetics. Biochemistry. https://doi.org/10.1021/bi960008e (1996).

Larnkjaer, A., Nykjaer, A., Olivecrona, G., Thøgersen, H. & Ostergaard, P. B. Structure of heparin fragments with high affinity for lipoprotein lipase and inhibition of lipoprotein lipase binding to alpha 2-macroglobulin-receptor/low-density-lipoprotein-receptor-related protein by heparin fragments. Biochem. J. https://doi.org/10.1042/bj3070205 (1995).

Spillmann, D., Lookene, A. & Olivecrona, G. Isolation and characterization of low sulfated heparan sulfate sequences with affinity for lipoprotein lipase. J. Biol. Chem. https://doi.org/10.1074/jbc.M604702200 (2006).

Young, E., Cosmi, B., Weitz, J. & Hirsh, J. Comparison of the non-specific binding of unfractionated heparin and low molecular weight heparin (Enoxaparin) to plasma proteins. Thromb Haemost. 70 (4), 625–630 (1993) ( PubMed PMID: 8115988 ).

Article   CAS   PubMed   Google Scholar  

Killeen, R., Wait, R., Begum, S., Gray, E. & Mulloy, B. Identification of major heparin-binding proteins in plasma using electrophoresis and mass spectrometry. Int. J. Exp. Pathol. https://doi.org/10.1111/j.0959-9673.2004.390af.x (2004).

Article   PubMed Central   Google Scholar  

Zammit, A., Pepper, D. S. & Dawes, J. Interaction of immobilised unfractionated and LMW heparins with proteins in whole human plasma. Thromb Haemost. 70 (6), 951–958 (1993) ( PubMed PMID: 8165617 ).

Xu, D. & Esko, J. D. Demystifying heparan sulfate-protein interactions. Annu. Rev. Biochem. https://doi.org/10.1146/annurev-biochem-060713-035314 (2014).

Bengtsson-Olivecrona, G. & Olivecrona, T. Phospholipase activity of milk lipoprotein lipase. Method. Enzymol. 197 , 345–356 (1991).

Article   CAS   Google Scholar  

Gill, S. & von Hippel, P. Calculation of protein extinction coefficients from amino acid sequence data. Anal. Biochem. https://doi.org/10.1016/0003-2697(89)90602-7 (1989).

Damen, J., Dijkstra, J., Regts, J. & Scherphof, G. Effect of lipoprotein-free plasma on the interaction of human plasma high density lipoprotein with egg yolk phosphatidylcholine liposomes. Biochimica et Biophysica Acta (BBa) - Lipids and Lipid Metabolism. https://doi.org/10.1016/0005-2760(80)90188-5 (1980).

Gunn, K. H., Neher, S. B., Gunn, K. H. & Neher, S. B. Structure of dimeric lipoprotein lipase reveals a pore adjacent to the active site. Nat. Commun. https://doi.org/10.1038/s41467-023-38243-9 (2023).

Mirdita, M. et al. ColabFold: making protein folding accessible to all. Nat. Methods. https://doi.org/10.1038/s41592-022-01488-1 (2022).

Evans, R. et al. Protein complex prediction with AlphaFold-Multimer. bioRxiv. https://doi.org/10.1101/2021.10.04.463034 (2022).

Consortium TU et al. UniProt: The universal protein knowledgebase in 2023. Nucleic Acids Res. https://doi.org/10.1093/nar/gkac1052 (2023).

Article   Google Scholar  

Meng, E. C. et al. UCSF ChimeraX: Tools for structure building and analysis. Protein Sci. https://doi.org/10.1002/pro.4792 (2023).

Risti, R. et al. Combined action of albumin and heparin regulates lipoprotein lipase oligomerization, stability, and ligand interactions. PLOS ONE. https://doi.org/10.1371/journal.pone.0283358 (2023).

Leth-Espensen, K. Z. et al. The intrinsic instability of the hydrolase domain of lipoprotein lipase facilitates its inactivation by ANGPTL4-catalyzed unfolding. Proc. Natl. Acad. Sci. https://doi.org/10.1073/pnas.2026650118 (2021).

Necci, M., Piovesan, D., Predictors, C., Curators, D. & Tosatto, S. C. E. AlphaFold and implications for intrinsically disordered proteins. J. Mol. Biol. https://doi.org/10.1016/j.jmb.2021.167208 (2021).

Doolittle, R. F. Biosynthesis Metabolism, Alterations in Disease. In The Plasma Proteins 2nd edn, Vol. II (ed. Putnam, F. W.) 148–9 (Academic Press, New York, 1975).

Google Scholar  

Olson, S. T., Halvorson, H. R. & Björk, I. Quantitative characterization of the thrombin-heparin interaction. Discrimination between specific and nonspecific binding models. J. Biol. Chem. https://doi.org/10.1016/S0021-9258(18)38124-9 (1991).

Du, X. et al. Insights into protein-ligand interactions: Mechanisms, models, and methods. Int. J. Mol. Sci. https://doi.org/10.3390/ijms17020144 (2016).

Wallerstein, J. et al. Entropy-entropy compensation between the protein, ligand, and solvent degrees of freedom fine-tunes affinity in ligand binding to galectin-3C. JACS Au. https://doi.org/10.1021/jacsau.0c00094 (2021).

Download references

Acknowledgements

This work was supported by Tallinn University of Technology (Grant SS22005 to A.L). The authors thank Jette Rindesalu for her contribution to several experiments.

Author information

Authors and affiliations.

Department of Chemistry and Biotechnology, Tallinn University of Technology, 12618, Tallinn, Estonia

Robert Risti, Mart Reimund, Natjan-Naatan Seeba & Aivar Lõokene

You can also search for this author in PubMed   Google Scholar

Contributions

A.L. and M.R. conceptualized the study. M.R. performed the experiments shown in Figs. 1 , 2 , 3 and 4 . R.R. compiled the data for Fig.  5 . R.R and N.-N.S. performed the experiments shown in Fig.  6 . R.R. performed the structural modeling on Fig.  7 . A.L. acquired funding for the project and supervised the project. R.R, M.R, A.L. wrote the main manuscript text. All authors reviewed the manuscript.

Corresponding author

Correspondence to Aivar Lõokene .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Supplementary information., rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/ .

Reprints and permissions

About this article

Cite this article.

Risti, R., Reimund, M., Seeba, NN. et al. A negatively charged cluster in the disordered acidic domain of GPIHBP1 provides selectivity in the interaction with lipoprotein lipase. Sci Rep 14 , 19639 (2024). https://doi.org/10.1038/s41598-024-70468-6

Download citation

Received : 30 April 2024

Accepted : 16 August 2024

Published : 23 August 2024

DOI : https://doi.org/10.1038/s41598-024-70468-6

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

lipoprotein lipase experiment

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

ijms-logo

Article Menu

lipoprotein lipase experiment

  • Subscribe SciFeed
  • Recommended Articles
  • Author Biographies
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

The role of selected lncrnas in lipid metabolism and cardiovascular disease risk, 1. introduction, 2. long noncoding rnas, 3. high-density lipoproteins, reverse cholesterol transport and cardiovascular risk, 4. lncrnas affecting lipoproteins, lipid metabolism, and atherosclerosis risk, 4.1. lncrna with beneficial effects, 4.1.1. nfia antisense rna 1.

  • Name/organism: RP5-833A20.1/ NFIA antisense RNA 1/NFIA-AS1; Homo sapiens
  • Databases &Code: Ensembl ENSG00000237853; HGNC: 40402; NCBI Gene: 645030
  • Chromosomal location& size: 1p31.3: 61,248,945–61,253,510 reverse strand; 4 exons, is associated with 1718 variant alleles
  • Interaction with other molecules: hsa-miR-382-5p
  • Biological consequence of such interaction: Increases circulation of HDL-C, reduces levels of LDL-C, and VLDL-C

4.1.2. Liver-Expressed LXR-Induced Sequence

  • Name/organism: LeXis/CT70 (cancer/testis associated transcript 70); Homo sapiens
  • Databases &Code: Ensembl ENSG00000230013; HGNC:37195
  • Chromosomal location& size: 9q31.1
  • Interaction with other molecules: Raly
  • Biological consequence of such interaction: Modulates expression of cholesterol biosynthetic genes

4.1.3. Macrophage-Expressed LXR-Induced Sequence

  • Name/organism: MeXis/AI427809/LOC381524; Homo sapiens
  • Databases &Code: Ensembl ENSMUSG00000086712.3
  • Chromosomal location& size: Chromosome 4: 53,261,356–53,270,232 reverse strand.
  • 4 transcripts (splice variants), 3 orthologues and is associated with 4 phenotypes
  • Interaction with other molecules: Raly, DDX17
  • Biological consequence of such interaction: Macrophage expressed LXRa (NR1H3)-dependent amplifier of Abca1 transcription lncRNA

4.1.4. LncRNA RP1-13D10.2

  • Name/organism: RP1-13D10.2;
  • Interaction with other molecules: LXR, SREBF2
  • Biological consequence of such interaction: Regulates LDLR gene expression in a sterol-responsive and SNP genotype–dependent manner in vitro

4.1.5. LncLSTR

  • Name/organism: LncLSTR (lncRNA liver-specific triglyceride regulator);
  • Chromosomal location& size: Syntenic to human chromosome 1q25
  • Interaction with other molecules: TDP-43
  • Biological consequence of such interaction: Modulates bile acid composition to regulate APOC2 expression, via FXR,85 and to control serum triglyceride levels

4.1.6. Cholesterol-Induced Regulator of Metabolism RNA

  • Name/organism: CHROME, PRKRA-AS1; Homo sapiens
  • Databases &Code: Ensemble: ENSG00000223960.9; HGNC:54059
  • Chromosomal location& size: 2q31.2: 178,413,635–178,440,243 forward strand; 35 transcripts (splice variants)
  • Interaction with other molecules: miR-27b, miR-33a, miR-33b, and miR-128
  • Biological consequence of such interaction: Promotes cholesterol secretion and HDL synthesis via suppressing the activity of specific miRNAs

4.1.7. Lipid-Droplet Transporter

  • Name/organism: LIPTER/LINC00881; Homo sapiens
  • Databases &Code: Ensemble: ENSG00000241135.8; HGNC: 48567; NCBI Gene: 100498859
  • Chromosomal location& size: 3q25.31: 157,089,634–157,135,557 forward strand; 11 transcripts
  • Interaction with other molecules: Phosphatidic acid, phosphatidylin-ositol 4-phosphate MYH10 motor protein
  • Biological consequence of such interaction: Facilitates the connection between LDs and the cytoskeleton for intracellular transport

4.1.8. Regulator of Hyperlipidaemia Long Noncoding RNA

  • Name/organism: lncRHPL; Mus musculus (house mouse)
  • Databases &Code: NIH Gene ID: 105244982
  • Chromosomal location& size: Chromosome 8
  • Interaction with other molecules: hnRNPU, BMAL1
  • Biological consequence of such interaction: Modulates hepatic VLDL secretion.

4.1.9. LncNONMUG027912

  • Name/organism: LncNONMMUG027912/lnc027912
  • Interaction with other molecules: AMPKα/mTOR signalling axis
  • Biological consequence of such interaction: Upregulates p-AMPKα, reduces p-mTOR levels, suppresses nuclear expression of SREBP1C, and hinders the expression of lipid synthesis genes

4.1.10. Maternally Expressed 3

  • Name/organism: MEG3/GTL2/LINC00023/NCRNA00023/ONCO-LNCRNA-83;
  • Databases &Code: ENSG00000214548.18; HGNC (14575); NCBI Gene (55384); OMIM ® (605636); Open Targets Plat-form (ENSG00000214548)
  • Chromosomal location& size: Chromosome 14: 100,779,410–100,861,031 forward strand;
  • 50 transcripts and is associated with 6 phenotypes
  • Interaction with other molecules: miR-21
  • Biological consequence of such interaction: Modulates hepatic lipogenesis

4.1.11. Cyclin-Dependent Kinase Inhibitor 2B Antisense RNA 1

  • Name/organism: ANRIL/CDKN2B-AS1/RP11-145E5.4/NCRNA00089/p15AS/PCAT12; Homo sapiens
  • Databases &Code: HGNC: 34341; Ensembl: ENSG00000240498; NCBI Gene: 100048912; OMIM ® : 613149
  • Chromosomal location& size: located within the CDKN2B-CDKN2A gene cluster at chromosome 9p21.3: 21,994,139–22,128,103 forward strand; 28 transcripts
  • Interaction with other molecules: polycomb repressive complex-1 (PRC1) and -2 (PRC2),
  • Biological consequence of such interaction: Epigenetic silencing of other genes in this cluster.

4.1.12. HOXC Cluster Antisense RNA 1

  • Name/organism: HOXC-AS1/NONHSAG011268.2/HSALNG0091321; Homo sapiens
  • Databases &Code: HGNC: 43749; NCBI Gene: 100874363; Ensembl: ENSG00000250451
  • Chromosomal location& size: 12q13.13; Ch 12: 53,999,022–54,000,010 reverse strand; 2 transcripts
  • Biological consequence of such interaction: Inhibition of intracellular lipid accumulation

4.2. LncRNA with Adverse Effects

4.2.1. ac068234.2–202 and ap001033.3–201.

  • Name/organism: AC068234.2–202; Homo sapiens
  • Databases &Code: AC068234.2
  • Chromosomal location& size: Ch17:47,303,474–47,323,613 reverse strand; transcript with 3 exons, associated with 4518 variant alleles
  • Interaction with other molecules: TBXA2R
  • Biological consequence of such interaction: Possibly contribute to the trans-regulation of the protein-coding gene thromboxane A2 receptor (TBXA2R)
  • Name/organism: AP001033.3–201; Homo sapiens
  • Databases &Code: AP001033.3
  • Chromosomal location& size: Ch18: 9,310,522–9,334,445 reverse strand; transcript with 3 exons and 5282 reported variant alleles
  • Interaction with other molecules: antisense to ITGB3
  • Biological consequence of such interaction: Acts a cis-regulator of the protein-coding gene integrin subunit beta 3 (ITGB3)

4.2.2. LncRNA ENST00000602558.1

  • Name/organism: ENST00000602558.1; Homo sapiens
  • Databases &Code: Ensembl: ENST00000602558.1
  • Chromosomal location& size: Chromosome 12: 123,971,457-123,971,714 reverse strand;
  • Exons: 1, Coding exons: 0, Transcript length: 258 bps; sense intronic to CCDC92
  • Interaction with other molecules: p65
  • Biological consequence of such interaction: Downregulates ABCG1 mRNA

4.2.3. Long Intergenic Non-Protein Coding RNA 1228

  • Name/organism: LINCRNA-DYNLRB2-2/LINC01228;
  • Databases &Code: Ensembl: ENST00000567966.1
  • Chromosomal location& size: Chromosome 16: 79,798,050–79,827,150 reverse strand; Size: 623 bp
  • Interaction with other molecules: GPR119
  • Biological consequence of such interaction: Facilitates cholesterol efflux and diminishes neutral lipid accumulation

4.2.4. Taurine Upregulated Gene 1

  • Name/organism: TUG1/FLJ20618/LINC00080/NCRNA00080; Homo sapiens
  • Databases &Code: Ensembl: ENSG00000253352.10
  • Chromosomal location& size: 22q12.2: 30,969,245–30,979,395 forward strand; 20 transcripts (splice variants) and 9 orthologues
  • Interaction with other molecules: miR-92a, miR-133a
  • Biological consequence of such interaction: Suppression of FGF1activation

4.2.5. Myocardial Infarction-Associated Transcript

  • Name/organism: MIAT/RNCR2/GOMAFU/C22orf35/LINC00066/NCRNA00066/lncRNA-MIAT; Homo sapiens
  • Databases &Code: HGNC: 33425; NCBI Gene: 440823; Ensembl: ENSG00000225783; OMIM ® : 611082
  • Chromosomal location& size: 22q12.1: 26,646,411–26,676,475 forward strand; 30 transcripts (splice variants) and is associated with 1 phenotype
  • Interaction with other molecules: PI3K/Akt signalling pathway
  • Biological consequence of such interaction: May constitute a component of the nuclear matrix; enhances angiogenesis and increases the expression of inflammatory factors

4.2.6. LncRNA RP11-728F11

  • Name/organism: LncRNA RP11-728F11;
  • Interaction with other molecules: EWSR1 (Ewings sarcoma RNA binding protein-1)
  • Biological consequence of such interaction: Induction of cholesterol uptake in monocytes-derived macrophages and proinflammatory cytokine production

4.2.7. lncRNA ENST00000416361

  • Name/organism: ENST00000416361; Homo sapiens
  • Databases &Code: Ensembl: ENST00000416361
  • Chromosomal location& size: 2102 bp
  • Interaction with other molecules: SREBP
  • Biological consequence of such interaction: Affects the occurrence and development of CAD

4.2.8. LncRNA RAPIA

  • Name/organism: RAPIA;
  • Chromosomal location& size: 10,252 nucleotides
  • Interaction with other molecules: miR-183-5p-ItgB1 (integrin β1)
  • Biological consequence of such interaction: Coordination of proliferation and apoptosis of macrophages

4.2.9. Nuclear Paraspeckle Assembly Transcript 1

  • Name/organism: NEAT1/LINC00084/MENEPSILON/BETA/NCRNA00084/TNCRNA/TP53LC15/VINC; Homo sapiens
  • Databases &Code: HGNC: 30815; NCBI Gene: 283131; Ensembl: ENSG00000245532; OMIM ® : 612769
  • Chromosomal location& size: 11q13.1; 11: 65,422,774–65,445,540 forward strand; 9 transcripts
  • Interaction with other molecules: miR-342-3p
  • Biological consequence of such interaction: Regulation of lipid droplet aggregation; affect TG metabolism

4.2.10. Nipsnap Homolog 3B

  • Name/organism: NIPSNAP3B/FP944/LOC286367/FLJ11275, SNAP1
  • Databases &Code: HGNC: 23641, NCBI Gene: 55335; Ensembl: ENSG00000165028; OMIM ® : 608872; UniProtKB/Swiss-Prot: Q9BS92
  • Chromosomal location& size: 9q31.1; Ch9: 104,764,129–104,777,764 forward strand; 3 transcripts (splice variants), 160 orthologues and 3 paralogues
  • Biological consequence of such interaction: Putative role in vesicular trafficking; promotion of intracellular lipid accumulation

4.2.11. Long Noncoding RNA Regulator of Akt Signalling Associated with HCC and RCC

  • Name/organism: LNCARSR/ lnc-TALC
  • Databases &Code: HGNC: 53864; NCBI Gene: 102723932; Ensembl: ENSG00000233086
  • Chromosomal location& size: 9q21.31; Ch 9: 79,505,804–79,567,802 reverse strand; 10 transcripts (splice variants)
  • Interaction with other molecules: SREBP-2
  • Biological consequence of such interaction: Promotion of the expression of HMG-CoA reductase (HMGCR), enhancement of hepatic de novo cholesterol synthesis rate

4.2.12. LDLR Antisense RNA 1

  • Name/organism: BM450697/LDLR-AS1
  • Databases &Code: HGNC: 54407, NCBI Gene: 115271120
  • Chromosomal location& size: 19p13.2; overlaps the 5′ UTR and coding sequence of the LDLR n the antisense orientation
  • Interaction with other molecules: PolII and potentially SREBP1a
  • Biological consequence of such interaction: Downregulation of the production of the low density lipoprotein receptor.

4.2.13. Long Non-Coding RNA Growth Arrest-Specific 5

  • Name/organism: GAS5/NCRNA00030/SNHG2; Homo sapiens
  • Databases &Code: HGNC: 16355; NCBI Gene: 60674; Ensembl: ENSG00000234741; OMIM ® : 608280
  • Chromosomal location& size: 1q25.1; Ch 1: 173,858,559–173,868,882 reverse strand; 91 transcripts (splice variants)
  • Interaction with other molecules: bind to the DNA binding domain of the glucocorticoid receptor (nuclear receptor subfamily 3, group C, member 1)
  • Biological consequence of such interaction: blockage of the activation of glucocorticoid receptor, regulation of the transcriptional activity of other receptors, such as androgen, progesterone and mineralocorticoid receptors

4.3. LncRNA with Ambiguous Effects

4.3.1. apolipoprotein a1 and a4 antisense rnas.

  • Name/organism: ApoA1-AS; Homo sapiens
  • Databases &Code: GeneCaRNA, HGNC: 40079, NCBI Gene: 104326055, Ensembl: ENSG00000235910, OMIM ® : 620112
  • Chromosomal location& size: 11q23.3, Size: 20,898 bases, Orientation: Plus strand
  • Interaction with other molecules: SUZ12, a component of the polycomb repressive complex 2 (PRC2)
  • Biological consequence of such interaction: Suppression of APOA1 expression
  • Name/organism: ApoA4-AS; mouse
  • Databases &Code: Ensembl and UCSC Genome Database
  • Chromosomal location& size: ∼900-nt
  • Interaction with other molecules: APOA4
  • Biological consequence of such interaction: APOA4-AS may regulate the expression of APOA4

4.3.2. lncRNA Induced by HCV, Regulator of SREBF1

  • Name/organism: LNCHR1; Homo sapiens
  • Databases &Code: Ensemble: ENSG00000257400.1; HGNC:56254
  • Chromosomal location& size: 12q22: 94,491,546–94,496,442 reverse strand; Size: 420 bp
  • Interaction with other molecules: SREB-1c
  • Biological consequence of such interaction: Regulation of the expression of SREBP-1-responsive genes

4.3.3. Solute Carrier Family 25 Member 15 (SLC25A15/lnc-HC)

  • Name/organism: lnc-HC/SLC25A15/HHH/ORC1/ORNT1/D13S327
  • Databases &Code: GenBank: MN026163.1
  • Chromosomal location& size: 1063 bp, linear
  • Interaction with other molecules: Coregulator: hnRNPA2B1
  • Biological consequence of such interaction: Reduction of the stability of mRNAs encoding Cyp7a1 and Abca1 (critical enzymes that contribute to cholesterol catabolism).

4.3.4. Metastasis-Associated Lung Adenocarcinoma Transcript 1

  • Name/organism: MALAT1; HCN, LINC00047, MASCRNA, NCRNA00047, NEAT2, PRO1073; Homo sapiens
  • Databases &Code: Ensembl: ENSG00000251562.11; HGNC:29665
  • Chromosomal location& size: 11q13.1: 65,497,640–65,508,073 forward strand; 66 transcripts (splice variants)
  • Interaction with other molecules: miR-17-ABCA1; miRNA-124-3p (sponge)
  • Biological consequence of such interaction: Contribution to cholesterol efflux, promotion of the upregulation of inflammatory CRP, modulation of PPARα expression.

4.3.5. A Novel Long Non-Coding RNA in Lipid Associated Single Nucleotide Polymorphism Gene Region

  • Name/organism: LASER/ LINC02702; Homo sapiens
  • Databases &Code: HGNC:54217; Ensembl: ENSG00000237937; NCBI Gene: 101929011
  • Chromosomal location& size: 11q23.3; 11: 116,639,422–116,658,295 forward strand; 4 transcripts
  • Interaction with other molecules: probably PCSK9
  • Biological consequence of such interaction: Enhancement of the expression of cholesterol metabolism genes

5. LncRNAs as Diagnostic and Therapeutic Targets in Lipid Disorders

6. conclusions, institutional review board statement, informed consent statement, data availability statement, conflicts of interest.

  • Zhang, X.; Price, N.L.; Fernández-Hernando, C. Non-coding RNAs in lipid metabolism. Vasc. Pharmacol. 2019 , 114 , 93–102. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • WHO. Cardiovascular Diseases. Available online: https://www.who.int/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds) (accessed on 15 December 2023).
  • Singh, D.D.; Kim, Y.; Choi, S.A.; Han, I.; Yadav, D.K. Clinical Significance of MicroRNAs, Long Non-Coding RNAs, and CircRNAs in Cardiovascular Diseases. Cells 2023 , 12 , 1629. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Poller, W.; Dimmeler, S.; Heymans, S.; Zeller, T.; Haas, J.; Karakas, M.; Leistner, D.-M.; Jakob, P.; Nakagawa, S.; Blankenberg, S. Non-coding RNAs in cardiovascular diseases: Diagnostic and therapeutic perspectives. Eur. Heart J. 2018 , 39 , 2704–2716. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Churov, A.; Summerhill, V.; Grechko, A.; Orekhova, V.; Orekhov, A. MicroRNAs as potential biomarkers in atherosclerosis. Int. J. Mol. Sci. 2019 , 20 , 5547. [ Google Scholar ] [ CrossRef ]
  • Collins, L.; Binder, P.; Chen, H.; Wang, X. Regulation of Long Non-coding RNAs and MicroRNAs in Heart Disease: Insight Into Mechanisms and Therapeutic Approaches. Front. Physiol. 2020 , 11 , 798. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yoon, J.-H.; Abdelmohsen, K.; Srikantan, S.; Yang, X.; Martindale, J.L.; De, S.; Huarte, M.; Zhan, M.; Becker, K.G.; Gorospe, M. LincRNA-p21 suppresses target mRNA translation. Mol. Cell 2012 , 47 , 648–655. [ Google Scholar ] [ CrossRef ]
  • Gong, C.; Maquat, L.E. lncRNAs transactivate STAU1-mediated mRNA decay by duplexing with 3′ UTRs via Alu elements. Nature 2011 , 470 , 284–288. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yin, J.; Zeng, X.; Ai, Z.; Yu, M.; Wu, Y.o.; Li, S. Construction and analysis of a lncRNA-miRNA-mRNA network based on competitive endogenous RNA reveal functional lncRNAs in oral cancer. BMC Med. Genom. 2020 , 13 , 84. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Paneru, B.; Ali, A.; Al-Tobasei, R.; Kenney, B.; Salem, M. Crosstalk among lncRNAs, microRNAs and mRNAs in the muscle ‘degradome’ of rainbow trout. Sci. Rep. 2018 , 8 , 8416. [ Google Scholar ] [ CrossRef ]
  • Mercer, T.R.; Mattick, J.S. Structure and function of long noncoding RNAs in epigenetic regulation. Nat. Struct. Mol. Biol. 2013 , 20 , 300–307. [ Google Scholar ] [ CrossRef ]
  • Das, S.; Shah, R.; Dimmeler, S.; Freedman, J.E.; Holley, C.; Lee, J.-M.; Moore, K.; Musunuru, K.; Wang, D.-Z.; Xiao, J.; et al. Noncoding RNAs in Cardiovascular Disease: Current Knowledge, Tools and Technologies for Investigation, and Future Directions: A Scientific Statement From the American Heart Association. Circ. Genom. Precis. Med. 2020 , 13 , e000062. [ Google Scholar ] [ CrossRef ]
  • Harrow, J.; Frankish, A.; Gonzalez, J.M.; Tapanari, E.; Diekhans, M.; Kokocinski, F.; Aken, B.L.; Barrell, D.; Zadissa, A.; Searle, S. GENCODE: The reference human genome annotation for The ENCODE Project. Genome Res. 2012 , 22 , 1760–1774. [ Google Scholar ] [ CrossRef ]
  • Zhao, Y.; Li, H.; Fang, S.; Kang, Y.; Wu, W.; Hao, Y.; Li, Z.; Bu, D.; Sun, N.; Zhang, M.Q. NONCODE 2016: An informative and valuable data source of long non-coding RNAs. Nucleic Acids Res. 2016 , 44 , D203–D208. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kung, J.T.; Colognori, D.; Lee, J.T. Long noncoding RNAs: Past, present, and future. Genetics 2013 , 193 , 651–669. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Aryal, B.; Rotllan, N.; Fernández-Hernando, C. Noncoding RNAs and atherosclerosis. Curr. Atheroscler. Rep. 2014 , 16 , 1–11. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ma, L.; Bajic, V.B.; Zhang, Z. On the classification of long non-coding RNAs. RNA Biol. 2013 , 10 , 924–933. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Schmitz, S.U.; Grote, P.; Herrmann, B.G. Mechanisms of long noncoding RNA function in development and disease. Cell. Mol. Life Sci. 2016 , 73 , 2491–2509. [ Google Scholar ] [ CrossRef ]
  • Wu, T.; Du, Y. LncRNAs: From basic research to medical application. Int. J. Biol. Sci. 2017 , 13 , 295. [ Google Scholar ] [ CrossRef ]
  • Palazzo, A.F.; Lee, E.S. Non-coding RNA: What is functional and what is junk? Front. Genet. 2015 , 6 , 2. [ Google Scholar ] [ CrossRef ]
  • Fang, Y.; Xu, Y.; Wang, R.; Hu, L.; Guo, D.; Xue, F.; Guo, W.; Zhang, D.; Hu, J.; Li, Y.; et al. Recent advances on the roles of LncRNAs in cardiovascular disease. J. Cell Mol. Med. 2020 , 24 , 12246–12257. [ Google Scholar ] [ CrossRef ]
  • Hon, C.-C.; Ramilowski, J.A.; Harshbarger, J.; Bertin, N.; Rackham, O.J.; Gough, J.; Denisenko, E.; Schmeier, S.; Poulsen, T.M.; Severin, J. An atlas of human long non-coding RNAs with accurate 5′ ends. Nature 2017 , 543 , 199–204. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liu, X.; Wang, T.-T.; Li, Y.; Shi, M.-M.; Li, H.-M.; Yuan, H.-X.; Mo, Z.-W.; Chen, J.; Zhang, B.; Chen, Y.-X.; et al. High density lipoprotein from coronary artery disease patients caused abnormal expression of long non-coding RNAs in vascular endothelial cells. Biochem. Biophys. Res. Commun. 2017 , 487 , 552–559. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zeng, Y.; Ren, K.; Zhu, X.; Zheng, Z.; Yi, G. Long Noncoding RNAs: Advances in Lipid Metabolism. Adv. Clin. Chem. 2018 , 87 , 1–36. [ Google Scholar ] [ CrossRef ]
  • Duan, J.; Huang, Z.; Nice, E.C.; Xie, N.; Chen, M.; Huang, C. Current advancements and future perspectives of long noncoding RNAs in lipid metabolism and signaling. J. Adv. Res. 2023 , 48 , 105–123. [ Google Scholar ] [ CrossRef ]
  • Wang, K.C.; Chang, H.Y. Molecular mechanisms of long noncoding RNAs. Mol. Cell 2011 , 43 , 904–914. [ Google Scholar ] [ CrossRef ]
  • Kino, T.; Hurt, D.E.; Ichijo, T.; Nader, N.; Chrousos, G.P. Noncoding RNA gas5 is a growth arrest–and starvation-associated repressor of the glucocorticoid receptor. Sci. Signal. 2010 , 3 , ra8. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ghafouri-Fard, S.; Dashti, S.; Taheri, M. The HOTTIP (HOXA transcript at the distal tip) lncRNA: Review of oncogenic roles in human. Biomed. Pharmacother. 2020 , 127 , 110158. [ Google Scholar ] [ CrossRef ]
  • Lee, J.T. The X as model for RNA’s niche in epigenomic regulation. Cold Spring Harb. Perspect. Biol. 2010 , 2 , a003749. [ Google Scholar ] [ CrossRef ]
  • Zhang, Y.; Du, W.; Yang, B. Long non-coding RNAs as new regulators of cardiac electrophysiology and arrhythmias: Molecular mechanisms, therapeutic implications and challenges. Pharmacol. Ther. 2019 , 203 , 107389. [ Google Scholar ] [ CrossRef ]
  • Sallam, T.; Sandhu, J.; Tontonoz, P. Long noncoding RNA discovery in cardiovascular disease: Decoding form to function. Circ. Res. 2018 , 122 , 155–166. [ Google Scholar ] [ CrossRef ]
  • Nelson, B.R.; Makarewich, C.A.; Anderson, D.M.; Winders, B.R.; Troupes, C.D.; Wu, F.; Reese, A.L.; McAnally, J.R.; Chen, X.; Kavalali, E.T. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science 2016 , 351 , 271–275. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, P.; Ruan, X.; Yang, L.; Kiesewetter, K.; Zhao, Y.; Luo, H.; Chen, Y.; Gucek, M.; Zhu, J.; Cao, H. A liver-enriched long non-coding RNA, lncLSTR, regulates systemic lipid metabolism in mice. Cell Metab. 2015 , 21 , 455–467. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, X.; Guo, S.; Hu, Y.; Guo, H.; Zhang, X.; Yan, Y.; Ma, J.; Li, Y.; Wang, H.; He, J.; et al. Microarray analysis of long non-coding RNA expression profiles in low high-density lipoprotein cholesterol disease. Lipids Health Dis. 2020 , 19 , 175. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Singh, A.K.; Aryal, B.; Zhang, X.; Fan, Y.; Price, N.L.; Suárez, Y.; Fernández-Hernando, C. Posttranscriptional regulation of lipid metabolism by non-coding RNAs and RNA binding proteins. In Seminars in Cell & Developmental Biology ; Academic Press: Cambridge, MA, USA, 2018; pp. 129–140. [ Google Scholar ]
  • Chen, Z. Progress and prospects of long noncoding RNAs in lipid homeostasis. Mol. Metab. 2016 , 5 , 164–170. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ross, R. Atherosclerosis—An inflammatory disease. N. Engl. J. Med. 1999 , 340 , 115–126. [ Google Scholar ] [ CrossRef ]
  • Glass, C.K.; Witztum, J.L. Atherosclerosis: The road ahead. Cell 2001 , 104 , 503–516. [ Google Scholar ] [ CrossRef ]
  • Rosenson, R.S.; Brewer Jr, H.B.; Davidson, W.S.; Fayad, Z.A.; Fuster, V.; Goldstein, J.; Hellerstein, M.; Jiang, X.-C.; Phillips, M.C.; Rader, D.J. Cholesterol efflux and atheroprotection: Advancing the concept of reverse cholesterol transport. Circulation 2012 , 125 , 1905–1919. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rader, D.J.; Hovingh, G.K. HDL and cardiovascular disease. Lancet 2014 , 384 , 618–625. [ Google Scholar ] [ CrossRef ]
  • Linton, M.F.; Tao, H.; Linton, E.F.; Yancey, P.G. SR-BI: A multifunctional receptor in cholesterol homeostasis and atherosclerosis. Trends Endocrinol. Metab. 2017 , 28 , 461–472. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dikkers, A.; Tietge, U.J. Biliary cholesterol secretion: More than a simple ABC. World J. Gastroenterol. WJG 2010 , 16 , 5936. [ Google Scholar ]
  • Nijstad, N.; Gautier, T.; Briand, F.; Rader, D.J.; Tietge, U.J. Biliary sterol secretion is required for functional in vivo reverse cholesterol transport in mice. Gastroenterology 2011 , 140 , 1043–1051. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Linsel-Nitschke, P.; Tall, A.R. HDL as a target in the treatment of atherosclerotic cardiovascular disease. Nat. Rev. Drug Discov. 2005 , 4 , 193–205. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Brewer, H.B. Focus on high-density lipoproteins in reducing cardiovascular risk. Am. Heart J. 2004 , 148 , S14–S18. [ Google Scholar ] [ CrossRef ]
  • Chang, F.-J.; Yuan, H.-Y.; Hu, X.-X.; Ou, Z.-J.; Fu, L.; Lin, Z.-B.; Wang, Z.-P.; Wang, S.-M.; Zhou, L.; Xu, Y.-Q.; et al. High density lipoprotein from patients with valvular heart disease uncouples endothelial nitric oxide synthase. J. Mol. Cell. Cardiol. 2014 , 74 , 209–219. [ Google Scholar ] [ CrossRef ]
  • Yuhanna, I.S.; Zhu, Y.; Cox, B.E.; Hahner, L.D.; Osborne-Lawrence, S.; Lu, P.; Marcel, Y.L.; Anderson, R.G.; Mendelsohn, M.E.; Hobbs, H.H.; et al. High-density lipoprotein binding to scavenger receptor-BI activates endothelial nitric oxide synthase. Nat. Med. 2001 , 7 , 853–857. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Barter, P.J.; Caulfield, M.; Eriksson, M.; Grundy, S.M.; Kastelein, J.J.P.; Komajda, M.; Lopez-Sendon, J.; Mosca, L.; Tardif, J.-C.; Waters, D.D.; et al. Effects of Torcetrapib in Patients at High Risk for Coronary Events. N. Engl. J. Med. 2007 , 357 , 2109–2122. [ Google Scholar ] [ CrossRef ]
  • Schwartz, G.G.; Olsson, A.G.; Abt, M.; Ballantyne, C.M.; Barter, P.J.; Brumm, J.; Chaitman, B.R.; Holme, I.M.; Kallend, D.; Leiter, L.A.; et al. Effects of Dalcetrapib in Patients with a Recent Acute Coronary Syndrome. N. Engl. J. Med. 2012 , 367 , 2089–2099. [ Google Scholar ] [ CrossRef ]
  • HPS2-THRIVE randomized placebo-controlled trial in 25,673 high-risk patients of ER niacin/laropiprant: Trial design, pre-specified muscle and liver outcomes, and reasons for stopping study treatment. Eur. Heart J. 2013 , 34 , 1279–1291. [ CrossRef ] [ PubMed ]
  • Franceschini, G.; Sirtori, C.R.; Capurso, A., II; Weisgraber, K.H.; Mahley, R.W. A-IMilano apoprotein. Decreased high density lipoprotein cholesterol levels with significant lipoprotein modifications and without clinical atherosclerosis in an Italian family. J. Clin. Investig. 1980 , 66 , 892–900. [ Google Scholar ] [ CrossRef ]
  • Speer, T.; Rohrer, L.; Blyszczuk, P.; Shroff, R.; Kuschnerus, K.; Kränkel, N.; Kania, G.; Zewinger, S.; Akhmedov, A.; Shi, Y.; et al. Abnormal high-density lipoprotein induces endothelial dysfunction via activation of Toll-like receptor-2. Immunity 2013 , 38 , 754–768. [ Google Scholar ] [ CrossRef ]
  • Sorrentino, S.A.; Besler, C.; Rohrer, L.; Meyer, M.; Heinrich, K.; Bahlmann, F.H.; Mueller, M.; Horváth, T.; Doerries, C.; Heinemann, M.; et al. Endothelial-vasoprotective effects of high-density lipoprotein are impaired in patients with type 2 diabetes mellitus but are improved after extended-release niacin therapy. Circulation 2010 , 121 , 110–122. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ma, X.-Y.; Liu, J.-P.; Song, Z.-Y. Associations of the ATP-binding cassette transporter A1 R219K polymorphism with HDL-C level and coronary artery disease risk: A meta-analysis. Atherosclerosis 2011 , 215 , 428–434. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, D.; Cheng, M.; Niu, Y.; Chi, X.; Liu, X.; Fan, J.; Fan, H.; Chang, Y.; Yang, W. Identification of a novel human long non-coding RNA that regulates hepatic lipid metabolism by inhibiting SREBP-1c. Int. J. Biol. Sci. 2017 , 13 , 349. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liu, C.; Yang, Z.; Wu, J.; Zhang, L.; Lee, S.; Shin, D.J.; Tran, M.; Wang, L. Long noncoding RNA H19 interacts with polypyrimidine tract-binding protein 1 to reprogram hepatic lipid homeostasis. Hepatology 2018 , 67 , 1768–1783. [ Google Scholar ] [ CrossRef ]
  • Kumar, S.; Williams, D.; Sur, S.; Wang, J.-Y.; Jo, H. Role of flow-sensitive microRNAs and long noncoding RNAs in vascular dysfunction and atherosclerosis. Vasc. Pharmacol. 2019 , 114 , 76–92. [ Google Scholar ] [ CrossRef ]
  • Hu, Y.-W.; Zhao, J.-Y.; Li, S.-F.; Huang, J.-L.; Qiu, Y.-R.; Ma, X.; Wu, S.-G.; Chen, Z.-P.; Hu, Y.-R.; Yang, J.-Y. RP5-833A20. 1/miR-382-5p/NFIA–dependent signal transduction pathway contributes to the regulation of cholesterol homeostasis and inflammatory reaction. Arterioscler. Thromb. Vasc. Biol. 2015 , 35 , 87–101. [ Google Scholar ] [ CrossRef ]
  • Wang, B.; Tontonoz, P. Liver X receptors in lipid signalling and membrane homeostasis. Nat. Rev. Endocrinol. 2018 , 14 , 452–463. [ Google Scholar ] [ CrossRef ]
  • Sallam, T.; Jones, M.C.; Gilliland, T.; Zhang, L.; Wu, X.; Eskin, A.; Sandhu, J.; Casero, D.; Vallim, T.Q.d.A.; Hong, C. Feedback modulation of cholesterol metabolism by the lipid-responsive non-coding RNA LeXis. Nature 2016 , 534 , 124–128. [ Google Scholar ] [ CrossRef ]
  • Sallam, T.; Jones, M.; Thomas, B.J.; Wu, X.; Gilliland, T.; Qian, K.; Eskin, A.; Casero, D.; Zhang, Z.; Sandhu, J. Transcriptional regulation of macrophage cholesterol efflux and atherogenesis by a long noncoding RNA. Nat. Med. 2018 , 24 , 304–312. [ Google Scholar ] [ CrossRef ]
  • Baggish, A.L.; Hale, A.; Weiner, R.B.; Lewis, G.D.; Systrom, D.; Wang, F.; Wang, T.J.; Chan, S.Y. Dynamic regulation of circulating microRNA during acute exhaustive exercise and sustained aerobic exercise training. J. Physiol. 2011 , 589 , 3983–3994. [ Google Scholar ] [ CrossRef ]
  • Mitchel, K.; Theusch, E.; Cubitt, C.; Dosé, A.C.; Stevens, K.; Naidoo, D.; Medina, M.W. RP1-13D10.2 Is a Novel Modulator of Statin-Induced Changes in Cholesterol. Circ. Cardiovasc. Genet. 2016 , 9 , 223–230. [ Google Scholar ] [ CrossRef ]
  • Zelcer, N.; Hong, C.; Boyadjian, R.; Tontonoz, P. LXR regulates cholesterol uptake through Idol-dependent ubiquitination of the LDL receptor. Science 2009 , 325 , 100–104. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • de Aguiar Vallim, T.Q.; Tarling, E.J.; Edwards, P.A. Pleiotropic roles of bile acids in metabolism. Cell Metab. 2013 , 17 , 657–669. [ Google Scholar ] [ CrossRef ]
  • Thomas, C.; Pellicciari, R.; Pruzanski, M.; Auwerx, J.; Schoonjans, K. Targeting bile-acid signalling for metabolic diseases. Nat. Rev. Drug Discov. 2008 , 7 , 678–693. [ Google Scholar ] [ CrossRef ]
  • Lee, E.B.; Lee, V.M.-Y.; Trojanowski, J.Q. Gains or losses: Molecular mechanisms of TDP43-mediated neurodegeneration. Nat. Rev. Neurosci. 2012 , 13 , 38–50. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ye, W.C.; Huang, S.F.; Hou, L.J.; Long, H.J.; Yin, K.; Hu, C.Y.; Zhao, G.J. Potential Therapeutic Targeting of lncRNAs in Cholesterol Homeostasis. Front. Cardiovasc. Med. 2021 , 8 , 688546. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ghafouri-Fard, S.; Gholipour, M.; Taheri, M. The Emerging Role of Long Non-coding RNAs and Circular RNAs in Coronary Artery Disease. Front. Cardiovasc. Med. 2021 , 8 , 632393. [ Google Scholar ] [ CrossRef ]
  • Hennessy, E.J.; van Solingen, C.; Scacalossi, K.R.; Ouimet, M.; Afonso, M.S.; Prins, J.; Koelwyn, G.J.; Sharma, M.; Ramkhelawon, B.; Carpenter, S.; et al. The long noncoding RNA CHROME regulates cholesterol homeostasis in primate. Nat. Metab. 2019 , 1 , 98–110. [ Google Scholar ] [ CrossRef ]
  • Han, L.; Huang, D.; Wu, S.; Liu, S.; Wang, C.; Sheng, Y.; Lu, X.; Broxmeyer, H.E.; Wan, J.; Yang, L. Lipid droplet-associated lncRNA LIPTER preserves cardiac lipid metabolism. Nat. Cell Biol. 2023 , 25 , 1033–1046. [ Google Scholar ] [ CrossRef ]
  • Olzmann, J.A.; Carvalho, P. Dynamics and functions of lipid droplets. Nat. Rev. Mol. Cell Biol. 2019 , 20 , 137–155. [ Google Scholar ] [ CrossRef ]
  • Goldberg, I.J.; Reue, K.; Abumrad, N.A.; Bickel, P.E.; Cohen, S.; Fisher, E.A.; Galis, Z.S.; Granneman, J.G.; Lewandowski, E.D.; Murphy, R. Deciphering the role of lipid droplets in cardiovascular disease: A report from the 2017 national heart, lung, and blood institute workshop. Circulation 2018 , 138 , 305–315. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Listenberger, L.L.; Han, X.; Lewis, S.E.; Cases, S.; Farese, R.V., Jr.; Ory, D.S.; Schaffer, J.E. Triglyceride accumulation protects against fatty acid-induced lipotoxicity. Proc. Natl. Acad. Sci. USA 2003 , 100 , 3077–3082. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Borradaile, N.M.; Schaffer, J.E. Lipotoxicity in the heart. Curr. Hypertens. Rep. 2005 , 7 , 412–417. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gao, Y.; Wang, Y.; Chen, X.; Peng, Y.; Chen, F.; He, Y.; Pang, W.; Yang, G.; Yu, T. MiR-127 attenuates adipogenesis by targeting MAPK4 and HOXC6 in porcine adipocytes. J. Cell. Physiol. 2019 , 234 , 21838–21850. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Geuens, T.; Bouhy, D.; Timmerman, V. The hnRNP family: Insights into their role in health and disease. Hum. Genet. 2016 , 135 , 851–867. [ Google Scholar ] [ CrossRef ]
  • Shen, X.; Zhang, Y.; Ji, X.; Li, B.; Wang, Y.; Huang, Y.; Zhang, X.; Yu, J.; Zou, R.; Qin, D.; et al. Long Noncoding RNA lncRHL Regulates Hepatic VLDL Secretion by Modulating hnRNPU/BMAL1/MTTP Axis. Diabetes 2022 , 71 , 1915–1928. [ Google Scholar ] [ CrossRef ]
  • Chu, K.; Zhao, N.; Hu, X.; Feng, R.; Zhang, L.; Wang, G.; Li, W.; Liu, L. LncNONMMUG027912 alleviates lipid accumulation through AMPKα/mTOR/SREBP1C axis in nonalcoholic fatty liver. Biochem. Biophys. Res. Commun. 2022 , 618 , 8–14. [ Google Scholar ] [ CrossRef ]
  • Liu, Y.; Jia, L.; Min, D.; Xu, Y.; Zhu, J.; Sun, Z. Baicalin inhibits proliferation and promotes apoptosis of vascular smooth muscle cells by regulating the MEG3/p53 pathway following treatment with ox-LDL. Int. J. Mol. Med. 2019 , 43 , 901–913. [ Google Scholar ] [ CrossRef ]
  • Yang, X.; Wang, C.C.; Lee, W.Y.W.; Trovik, J.; Chung, T.K.H.; Kwong, J. Long non-coding RNA HAND2-AS1 inhibits invasion and metastasis in endometrioid endometrial carcinoma through inactivating neuromedin U. Cancer Lett. 2018 , 413 , 23–34. [ Google Scholar ] [ CrossRef ]
  • Zhu, X.; Li, H.; Wu, Y.; Zhou, J.; Yang, G.; Wang, W. lncRNA MEG3 promotes hepatic insulin resistance by serving as a competing endogenous RNA of miR-214 to regulate ATF4 expression. Int. J. Mol. Med. 2019 , 43 , 345–357. [ Google Scholar ] [ CrossRef ]
  • Huang, P.; Huang, F.-Z.; Liu, H.-Z.; Zhang, T.-Y.; Yang, M.-S.; Sun, C.-Z. LncRNA MEG3 functions as a ceRNA in regulating hepatic lipogenesis by competitively binding to miR-21 with LRP6. Metabolism 2019 , 94 , 1–8. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, H.; Han, S.; Sun, Q.; Yao, Y.; Li, S.; Yuan, C.; Zhang, B.; Jing, B.; Wu, J.; Song, Y.; et al. Long non-coding RNA CDKN2B-AS1 reduces inflammatory response and promotes cholesterol efflux in atherosclerosis by inhibiting ADAM10 expression. Aging 2019 , 11 , 1695–1715. [ Google Scholar ] [ CrossRef ]
  • Huang, C.; Hu, Y.-W.; Zhao, J.-J.; Ma, X.; Zhang, Y.; Guo, F.-X.; Kang, C.-M.; Lu, J.-B.; Xiu, J.-C.; Sha, Y.-H. Long noncoding RNA HOXC-AS1 suppresses Ox-LDL-induced cholesterol accumulation through promoting HOXC6 expression in THP-1 macrophages. DNA Cell Biol. 2016 , 35 , 722–729. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Papp, E.; Havasi, V.; Bene, J.; Komlosi, K.; Czopf, L.; Magyar, E.; Feher, C.; Feher, G.; Horvath, B.; Marton, Z. Glycoprotein IIIA gene (PIA) polymorphism and aspirin resistance: Is there any correlation? Ann. Pharmacother. 2005 , 39 , 1013–1018. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Morgan, E.A.; Schneider, J.G.; Baroni, T.E.; Uluçkan, Ö.; Heller, E.; Hurchla, M.A.; Deng, H.; Floyd, D.; Berdy, A.; Prior, J.L. Dissection of platelet and myeloid cell defects by conditional targeting of the β3-integrin subunit. FASEB J. 2010 , 24 , 1117. [ Google Scholar ] [ CrossRef ]
  • Shao, J.; Fu, Y.; Yang, W.; Yan, J.; Zhao, J.; Chen, S.; Xia, W. Thromboxane A2 receptor polymorphism in association with cerebral infarction and its regulation on platelet function. Curr. Neurovasc. Res. 2015 , 12 , 15–24. [ Google Scholar ] [ CrossRef ]
  • Cai, C.; Zhu, H.; Ning, X.; Li, L.; Yang, B.; Chen, S.; Wang, L.; Lu, X.; Gu, D. LncRNA ENST00000602558.1 regulates ABCG1 expression and cholesterol efflux from vascular smooth muscle cells through a p65-dependent pathway. Atherosclerosis 2019 , 285 , 31–39. [ Google Scholar ] [ CrossRef ]
  • Ageeli Hakami, M. Diabetes and diabetic associative diseases: An overview of epigenetic regulations of TUG1. Saudi J. Biol. Sci. 2024 , 31 , 103976. [ Google Scholar ] [ CrossRef ]
  • Yang, L.; Li, T. LncRNA TUG1 regulates ApoM to promote atherosclerosis progression through miR-92a/FXR1 axis. J. Cell. Mol. Med. 2020 , 24 , 8836–8848. [ Google Scholar ] [ CrossRef ]
  • Zhang, L.; Cheng, H.; Yue, Y.; Li, S.; Zhang, D.; He, R. TUG1 knockdown ameliorates atherosclerosis via up-regulating the expression of miR-133a target gene FGF1. Cardiovasc. Pathol. 2018 , 33 , 6–15. [ Google Scholar ] [ CrossRef ]
  • Zhu, X.H.; Yuan, Y.X.; Rao, S.L.; Wang, P. LncRNA MIAT enhances cardiac hypertrophy partly through sponging miR-150. Eur. Rev. Med. Pharmacol. Sci. 2016 , 20 , 3653–3660. [ Google Scholar ] [ PubMed ]
  • Ye, Z.-M.; Yang, S.; Xia, Y.-P.; Hu, R.-T.; Chen, S.; Li, B.-W.; Chen, S.-L.; Luo, X.-Y.; Mao, L.; Li, Y. LncRNA MIAT sponges miR-149-5p to inhibit efferocytosis in advanced atherosclerosis through CD47 upregulation. Cell Death Dis. 2019 , 10 , 138. [ Google Scholar ] [ CrossRef ]
  • Sun, G.; Li, Y.; Ji, Z. Up-regulation of MIAT aggravates the atherosclerotic damage in atherosclerosis mice through the activation of PI3K/Akt signaling pathway. Drug Deliv. 2019 , 26 , 641–649. [ Google Scholar ] [ CrossRef ]
  • Li, H.; Zhang, Y.; Guo, Y.; Liu, R.; Yu, Q.; Gong, L.; Liu, Z.; Xie, W.; Wang, C. ALKBH1 promotes lung cancer by regulating m6A RNA demethylation. Biochem. Pharmacol. 2021 , 189 , 114284. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wu, L.; Pei, Y.; Zhu, Y.; Jiang, M.; Wang, C.; Cui, W.; Zhang, D. Association of N6-methyladenine DNA with plaque progression in atherosclerosis via myocardial infarction-associated transcripts. Cell Death Dis. 2019 , 10 , 909. [ Google Scholar ] [ CrossRef ]
  • Sun, C.; Huang, L.; Li, Z.; Leng, K.; Xu, Y.; Jiang, X.; Cui, Y. Long non-coding RNA MIAT in development and disease: A new player in an old game. J. Biomed. Sci. 2018 , 25 , 1–7. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yan, B.; Yao, J.; Liu, J.-Y.; Li, X.-M.; Wang, X.-Q.; Li, Y.-J.; Tao, Z.-F.; Song, Y.-C.; Chen, Q.; Jiang, Q. lncRNA-MIAT regulates microvascular dysfunction by functioning as a competing endogenous RNA. Circ. Res. 2015 , 116 , 1143–1156. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dong, X.H.; Lu, Z.F.; Kang, C.M.; Li, X.H.; Haworth, K.E.; Ma, X.; Lu, J.B.; Liu, X.H.; Fang, F.C.; Wang, C.S.; et al. The Long Noncoding RNA RP11-728F11.4 Promotes Atherosclerosis. Arter. Thromb. Vasc. Biol. 2021 , 41 , 1191–1204. [ Google Scholar ] [ CrossRef ]
  • Li, P.; Yan, X.; Xu, G.; Pang, Z.; Weng, J.; Yin, J.; Li, M.; Yu, L.; Chen, Q.; Sun, K. A novel plasma lncRNA ENST00000416361 is upregulated in coronary artery disease and is related to inflammation and lipid metabolism. Mol. Med. Rep. 2020 , 21 , 2375–2384. [ Google Scholar ] [ CrossRef ]
  • Brown, M.S.; Goldstein, J.L. The SREBP Pathway: Regulation of Cholesterol Metabolism by Proteolysis of a Membrane-Bound Transcription Factor. Cell 1997 , 89 , 331–340. [ Google Scholar ] [ CrossRef ]
  • Sato, R. Sterol metabolism and SREBP activation. Arch. Biochem. Biophys. 2010 , 501 , 177–181. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pérez-Belmonte, L.M.; Moreno-Santos, I.; Cabrera-Bueno, F.; Sánchez-Espín, G.; Castellano, D.; Such, M.; Crespo-Leiro, M.G.; Carrasco-Chinchilla, F.; Alonso-Pulpón, L.; López-Garrido, M.; et al. Expression of Sterol Regulatory Element-Binding Proteins in epicardial adipose tissue in patients with coronary artery disease and diabetes mellitus: Preliminary study. Int. J. Med. Sci. 2017 , 14 , 268–274. [ Google Scholar ] [ CrossRef ]
  • Sun, C.; Fu, Y.; Gu, X.; Xi, X.; Peng, X.; Wang, C.; Sun, Q.; Wang, X.; Qian, F.; Qin, Z.; et al. Macrophage-Enriched lncRNA RAPIA. Arterioscler. Thromb. Vasc. Biol. 2020 , 40 , 1464–1478. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hu, S.; Liu, Y.; You, T.; Heath, J.; Xu, L.; Zheng, X.; Wang, A.; Wang, Y.; Li, F.; Yang, F. Vascular semaphorin 7A upregulation by disturbed flow promotes atherosclerosis through endothelial β1 integrin. Arterioscler. Thromb. Vasc. Biol. 2018 , 38 , 335–343. [ Google Scholar ] [ CrossRef ]
  • Pan, Y.; Xin, W.; Wei, W.; Tatenhorst, L.; Graf, I.; Popa-Wagner, A.; Gerner, S.T.; Huber, S.E.; Kilic, E.; Hermann, D.M.; et al. Knockdown of NEAT1 prevents post-stroke lipid droplet agglomeration in microglia by regulating autophagy. Cell Mol. Life Sci. 2024 , 81 , 30. [ Google Scholar ] [ CrossRef ]
  • Vlachogiannis, N.I.; Sachse, M.; Georgiopoulos, G.; Zormpas, E.; Bampatsias, D.; Delialis, D.; Bonini, F.; Galyfos, G.; Sigala, F.; Stamatelopoulos, K. Adenosine-to-inosine Alu RNA editing controls the stability of the pro-inflammatory long noncoding RNA NEAT1 in atherosclerotic cardiovascular disease. J. Mol. Cell. Cardiol. 2021 , 160 , 111–120. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, L.; Xia, J.W.; Ke, Z.P.; Zhang, B.H. Blockade of NEAT1 represses inflammation response and lipid uptake via modulating miR-342-3p in human macrophages THP-1 cells. J. Cell. Physiol. 2019 , 234 , 5319–5326. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Du, X.J.; Wei, J.; Tian, D.; Yan, C.; Hu, P.; Wu, X.; Yang, W.; Hu, X. NEAT1 promotes myocardial ischemia-reperfusion injury via activating the MAPK signaling pathway. J. Cell. Physiol. 2019 , 234 , 18773–18780. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cui, N.; Hu, M.; Khalil, R.A. Biochemical and biological attributes of matrix metalloproteinases. Prog. Mol. Biol. Transl. Sci. 2017 , 147 , 1–73. [ Google Scholar ]
  • Luo, M.; Sun, Q.; Zhao, H.; Tao, J.; Yan, D. Long noncoding RNA NEAT1 sponges miR-495-3p to enhance myocardial ischemia-reperfusion injury via MAPK6 activation. J. Cell. Physiol. 2020 , 235 , 105–113. [ Google Scholar ] [ CrossRef ]
  • Fan, G.; Zhang, C.; Wei, X.; Wei, R.; Qi, Z.; Chen, K.; Cai, X.; Xu, L.; Tang, L.; Zhou, J. NEAT1/hsa-miR-372–3p axis participates in rapaMYCin-induced lipid metabolic disorder. Free Radic. Biol. Med. 2021 , 167 , 1–11. [ Google Scholar ] [ CrossRef ]
  • Liu, X.; Liang, Y.; Song, R.; Yang, G.; Han, J.; Lan, Y.; Pan, S.; Zhu, M.; Liu, Y.; Wang, Y. Long non-coding RNA NEAT1-modulated abnormal lipolysis via ATGL drives hepatocellular carcinoma proliferation. Mol. Cancer 2018 , 17 , 1–18. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ma, X.; Wang, T.; Zhao, Z.-L.; Jiang, Y.; Ye, S. Propofol suppresses proinflammatory cytokine production by increasing ABCA1 expression via mediation by the long noncoding RNA LOC286367. Mediat. Inflamm. 2018 , 2018 , 8907143. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Joyce, C.W.; Amar, M.J.; Lambert, G.; Vaisman, B.L.; Paigen, B.; Najib-Fruchart, J.; Hoyt Jr, R.F.; Neufeld, E.D.; Remaley, A.T.; Fredrickson, D.S. The ATP binding cassette transporter A1 (ABCA1) modulates the development of aortic atherosclerosis in C57BL/6 and apoE-knockout mice. Proc. Natl. Acad. Sci. USA 2002 , 99 , 407–412. [ Google Scholar ] [ CrossRef ]
  • Gao, J.-H.; He, L.-H.; Yu, X.-H.; Zhao, Z.-W.; Wang, G.; Zou, J.; Wen, F.-J.; Zhou, L.; Wan, X.-J.; Zhang, D.-W. CXCL12 promotes atherosclerosis by downregulating ABCA1 expression via the CXCR4/GSK3β/β-cateninT120/TCF21 pathway. J. Lipid Res. 2019 , 60 , 2020–2033. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Huang, J.; Chen, S.; Cai, D.; Bian, D.; Wang, F. Long noncoding RNA lncARSR promotes hepatic cholesterol biosynthesis via modulating Akt/SREBP-2/HMGCR pathway. Life Sci. 2018 , 203 , 48–53. [ Google Scholar ] [ CrossRef ]
  • Cirera-Salinas, D.; Pauta, M.; Allen, R.M.; Salerno, A.G.; Ramírez, C.M.; Chamorro-Jorganes, A.; Wanschel, A.C.; Lasuncion, M.A.; Morales-Ruiz, M.; Suarez, Y. Mir-33 regulates cell proliferation and cell cycle progression. Cell Cycle 2012 , 11 , 922–933. [ Google Scholar ] [ CrossRef ]
  • Go, G.-W.; Mani, A. Low-density lipoprotein receptor (LDLR) family orchestrates cholesterol homeostasis. Yale J. Biol. Med. 2012 , 85 , 19. [ Google Scholar ]
  • Ray, R.M.; Hansen, A.H.; Slott, S.; Taskova, M.; Astakhova, K.; Morris, K.V. Control of LDL uptake in human cells by targeting the LDLR regulatory long non-coding RNA BM450697. Mol. Ther. -Nucleic Acids 2019 , 17 , 264–276. [ Google Scholar ] [ CrossRef ]
  • Chen, L.; Yao, H.; Hui, J.Y.; Ding, S.H.; Fan, Y.L.; Pan, Y.H.; Chen, K.H.; Wan, J.Q.; Jiang, J.Y. Global transcriptomic study of atherosclerosis development in rats. Gene 2016 , 592 , 43–48. [ Google Scholar ] [ CrossRef ]
  • Meng, X.-D.; Yao, H.-H.; Wang, L.-M.; Yu, M.; Shi, S.; Yuan, Z.-X.; Liu, J. Knockdown of GAS5 inhibits atherosclerosis progression via reducing EZH2-mediated ABCA1 transcription in ApoE−/− mice. Mol. Ther.-Nucleic Acids 2020 , 19 , 84–96. [ Google Scholar ] [ CrossRef ]
  • Halley, P.; Kadakkuzha, B.M.; Faghihi, M.A.; Magistri, M.; Zeier, Z.; Khorkova, O.; Coito, C.; Hsiao, J.; Lawrence, M.; Wahlestedt, C. Regulation of the apolipoprotein gene cluster by a long noncoding RNA. Cell Rep. 2014 , 6 , 222–230. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Qin, W.; Li, X.; Xie, L.; Li, S.; Liu, J.; Jia, L.; Dong, X.; Ren, X.; Xiao, J.; Yang, C. A long non-coding RNA, APOA4-AS, regulates APOA4 expression depending on HuR in mice. Nucleic Acids Res. 2016 , 44 , 6423–6433. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lan, X.; Yan, J.; Ren, J.; Zhong, B.; Li, J.; Li, Y.; Liu, L.; Yi, J.; Sun, Q.; Yang, X. A novel long noncoding RNA Lnc-HC binds hnRNPA2B1 to regulate expressions of Cyp7a1 and Abca1 in hepatocytic cholesterol metabolism. Hepatology 2016 , 64 , 58–72. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lan, X.; Wu, L.; Wu, N.; Chen, Q.; Li, Y.; Du, X.; Wei, C.; Feng, L.; Li, Y.; Osoro, E.K. Long noncoding RNA lnc-HC regulates PPARγ-mediated hepatic lipid metabolism through miR-130b-3p. Mol. Ther.-Nucleic Acids 2019 , 18 , 954–965. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lv, F.; Liu, L.; Feng, Q.; Yang, X. Long non-coding RNA MALAT1 and its target microRNA-125b associate with disease risk, severity, and major adverse cardiovascular event of coronary heart disease. J. Clin. Lab. Anal. 2021 , 35 , e23593. [ Google Scholar ] [ CrossRef ]
  • Guo, Y.; Luo, F.; Liu, Q.; Xu, D. Regulatory non-coding RNAs in acute myocardial infarction. J. Cell. Mol. Med. 2017 , 21 , 1013–1023. [ Google Scholar ] [ CrossRef ]
  • Liu, L.; Tan, L.; Yao, J.; Yang, L. Long non-coding RNA MALAT1 regulates cholesterol accumulation in ox-LDL-induced macrophages via the microRNA-17-5p/ABCA1 axis. Mol. Med. Rep. 2020 , 21 , 1761–1770. [ Google Scholar ] [ CrossRef ]
  • Liu, Z.; Liu, J.; Wei, Y.; Xu, J.; Wang, Z.; Wang, P.; Sun, H.; Song, Z.; Liu, Q. LncRNA MALAT1 prevents the protective effects of miR-125b-5p against acute myocardial infarction through positive regulation of NLRC5. Exp. Ther. Med. 2020 , 19 , 990–998. [ Google Scholar ] [ CrossRef ]
  • Chao, C.T.; Yeh, H.Y.; Yuan, T.H.; Chiang, C.K.; Chen, H.W. MicroRNA-125b in vascular diseases: An updated systematic review of pathogenetic implications and clinical applications. J. Cell. Mol. Med. 2019 , 23 , 5884–5894. [ Google Scholar ] [ CrossRef ]
  • Ding, X.-Q.; Ge, P.-C.; Liu, Z.; Jia, H.; Chen, X.; An, F.-H.; Li, L.-H.; Chen, Z.-H.; Mao, H.-W.; Li, Z.-Y. Interaction between microRNA expression and classical risk factors in the risk of coronary heart disease. Sci. Rep. 2015 , 5 , 14925. [ Google Scholar ] [ CrossRef ]
  • Akhabue, E.; Thiboutot, J.; Cheng, J.-w.; Lerakis, S.; Vittorio, T.J.; Christodoulidis, G.; Grady, K.M.; Kosmas, C.E. New and emerging risk factors for coronary heart disease. Am. J. Med. Sci. 2014 , 347 , 151–158. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, C.; Hu, Z.; Zhang, W.; Yu, J.; Yang, Y.; Xu, Z.; Luo, H.; Liu, X.; Liu, Y.; Chen, C. Regulation of cholesterol homeostasis by a novel long non-coding RNA LASER. Sci. Rep. 2019 , 9 , 7693. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Seidah, N.G.; Awan, Z.; Chrétien, M.; Mbikay, M. PCSK9: A key modulator of cardiovascular health. Circ. Res. 2014 , 114 , 1022–1036. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lee, K.-H.; Hwang, H.-J.; Cho, J.-Y. Long Non-Coding RNA Associated with Cholesterol Homeostasis and Its Involvement in Metabolic Diseases. Int. J. Mol. Sci. 2020 , 21 , 8337. [ Google Scholar ] [ CrossRef ]
  • Paredes, S.; Fonseca, L.; Ribeiro, L.; Ramos, H.; Oliveira, J.C.; Palma, I. Novel and traditional lipid profiles in metabolic syndrome reveal a high atherogenicity. Sci. Rep. 2019 , 9 , 11792. [ Google Scholar ] [ CrossRef ]
  • Cox, R.; García-Palmieri, M. Cholesterol, Triglycerides, and Associated Lipoproteins. In Clinical Methods: The History, Physical, and Laboratory Examinations , 3rd ed.; Walker, H.K., Hall, W.D., Hurst, J.W., Eds.; Butterworths: Boston, MA, USA, 1990. [ Google Scholar ]
  • Wang, W.-T.; Sun, Y.-M.; Huang, W.; He, B.; Zhao, Y.-N.; Chen, Y.-Q. Genome-wide long non-coding RNA analysis identified circulating LncRNAs as novel non-invasive diagnostic biomarkers for gynecological disease. Sci. Rep. 2016 , 6 , 23343. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pardini, B.; Sabo, A.A.; Birolo, G.; Calin, G.A. Noncoding RNAs in extracellular fluids as cancer biomarkers: The new frontier of liquid biopsies. Cancers 2019 , 11 , 1170. [ Google Scholar ] [ CrossRef ]
  • Huang, S.F.; Peng, X.F.; Jiang, L.; Hu, C.Y.; Ye, W.C. LncRNAs as Therapeutic Targets and Potential Biomarkers for Lipid-Related Diseases. Front. Pharmacol. 2021 , 12 , 729745. [ Google Scholar ] [ CrossRef ]
  • Pan, J.-X. LncRNA H19 promotes atherosclerosis by regulating MAPK and NF-kB signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2017 , 21 , 322–328. [ Google Scholar ]
  • Tao, K.; Hu, Z.; Zhang, Y.; Jiang, D.; Cheng, H. LncRNA CASC11 improves atherosclerosis by downregulating IL-9 and regulating vascular smooth muscle cell apoptosis and proliferation. Biosci. Biotechnol. Biochem. 2019 , 83 , 1284–1288. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chen, L.; Yang, W.; Guo, Y.; Chen, W.; Zheng, P.; Zeng, J.; Tong, W. Exosomal lncRNA GAS5 regulates the apoptosis of macrophages and vascular endothelial cells in atherosclerosis. PLoS ONE 2017 , 12 , e0185406. [ Google Scholar ] [ CrossRef ]
  • Li, F.-P.; Lin, D.-Q.; Gao, L.-Y. LncRNA TUG1 promotes proliferation of vascular smooth muscle cell and atherosclerosis through regulating miRNA-21/PTEN axis. Eur. Rev. Med. Pharmacol. Sci. 2018 , 22 , 7439–7447. [ Google Scholar ] [ PubMed ]
  • Yao, X.; Yan, C.; Zhang, L.; Li, Y.; Wan, Q. LncRNA ENST00113 promotes proliferation, survival, and migration by activating PI3K/Akt/mTOR signaling pathway in atherosclerosis. Medicine 2018 , 97 , e0473. [ Google Scholar ] [ CrossRef ]
  • Hu, Y.-W.; Guo, F.-X.; Xu, Y.-J.; Li, P.; Lu, Z.-F.; McVey, D.G.; Zheng, L.; Wang, Q.; John, H.Y.; Kang, C.-M. Long noncoding RNA NEXN-AS1 mitigates atherosclerosis by regulating the actin-binding protein NEXN. J. Clin. Investig. 2019 , 129 , 1115–1128. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yao, Y.; Zhang, T.; Qi, L.; Zhou, C.; Wei, J.; Feng, F.; Liu, R.; Sun, C. Integrated analysis of co-expression and ceRNA network identifies five lncRNAs as prognostic markers for breast cancer. J. Cell. Mol. Med. 2019 , 23 , 8410–8419. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Park, J.G.; Kim, G.; Jang, S.Y.; Lee, Y.R.; Lee, E.; Lee, H.W.; Han, M.-H.; Chun, J.M.; Han, Y.S.; Yoon, J.S. Plasma long noncoding RNA LeXis is a potential diagnostic marker for non-alcoholic steatohepatitis. Life 2020 , 10 , 230. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mannu, G.; Zaman, M.; Gupta, A.; Rehman, H.; Myint, P. Evidence of lifestyle modification in the management of hypercholesterolemia. Curr. Cardiol. Rev. 2013 , 9 , 2–14. [ Google Scholar ]
  • Chi, X.; Gatti, P.; Papoian, T. Safety of antisense oligonucleotide and siRNA-based therapeutics. Drug Discov. Today 2017 , 22 , 823–833. [ Google Scholar ] [ CrossRef ]
  • Liu, S.J.; Horlbeck, M.A.; Cho, S.W.; Birk, H.S.; Malatesta, M.; He, D.; Attenello, F.J.; Villalta, J.E.; Cho, M.Y.; Chen, Y. CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells. Science 2017 , 355 , eaah7111. [ Google Scholar ] [ CrossRef ]
  • Zhang, H.; Zou, X.; Liu, F. Silencing TTTY15 mitigates hypoxia-induced mitochondrial energy metabolism dysfunction and cardiomyocytes apoptosis via TTTY15/let-7i-5p and TLR3/NF-κB pathways. Cell. Signal. 2020 , 76 , 109779. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gong, N.; Teng, X.; Li, J.; Liang, X.-J. Antisense oligonucleotide-conjugated nanostructure-targeting lncRNA MALAT1 inhibits cancer metastasis. ACS Appl. Mater. Interfaces 2018 , 11 , 37–42. [ Google Scholar ] [ CrossRef ]
  • Bernardo, B.C.; Ooi, J.Y.; Lin, R.C.; McMullen, J.R. miRNA therapeutics: A new class of drugs with potential therapeutic applications in the heart. Future Med. Chem. 2015 , 7 , 1771–1792. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lee, J.-S.; Mendell, J.T. Antisense-mediated transcript knockdown triggers premature transcription termination. Mol. Cell 2020 , 77 , 1044–1054.e1043. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wan, W.B.; Seth, P.P. The medicinal chemistry of therapeutic oligonucleotides. J. Med. Chem. 2016 , 59 , 9645–9667. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dhuri, K.; Bechtold, C.; Quijano, E.; Pham, H.; Gupta, A.; Vikram, A.; Bahal, R. Antisense oligonucleotides: An emerging area in drug discovery and development. J. Clin. Med. 2020 , 9 , 2004. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Stein, C.A.; Castanotto, D. FDA-approved oligonucleotide therapies in 2017. Mol. Ther. 2017 , 25 , 1069–1075. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Bhat, S.A.; Ahmad, S.M.; Mumtaz, P.T.; Malik, A.A.; Dar, M.A.; Urwat, U.; Shah, R.A.; Ganai, N.A. Long non-coding RNAs: Mechanism of action and functional utility. Non-Coding RNA Res. 2016 , 1 , 43–50. [ Google Scholar ] [ CrossRef ]
  • Bobbin, M.L.; Rossi, J.J. RNA interference (RNAi)-based therapeutics: Delivering on the promise? Annu. Rev. Pharmacol. Toxicol. 2016 , 56 , 103–122. [ Google Scholar ] [ CrossRef ]
  • Winter, H.; Winski, G.; Busch, A.; Chernogubova, E.; Fasolo, F.; Wu, Z.; Bäcklund, A.; Khomtchouk, B.B.; Van Booven, D.J.; Sachs, N.; et al. Targeting long non-coding RNA NUDT6 enhances smooth muscle cell survival and limits vascular disease progression. Mol. Ther. 2023 , 31 , 1775–1790. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

lncRNAImpact on Lipid Metabolism and Cardiovascular RiskRefs.
lncRNA H19 [ ]
AC068234.2–202 [ ]
AP001033.3–201 [ ]
ApoA1-AS [ ]
ApoA4-AS [ ]
Overexpression of ENST00000602558.1 [ ]
RP5-833A20.1 [ , ]
LeXis [ , ]
MeXis [ , , ]
LncHR1 [ ]
RP1-13D10.2 [ , ]
LncLSTR [ ]
Lnc-HC [ , , ]
LincRNA-DYNLRB2-2 [ ]
CHROME (PRKRA-AS1) [ , , ]
lncRNA LIPTER [ , ]
lncRHPL [ ]
LncNONMMUG027912 [ ]
MALAT1 [ , , , , , ]
TUG1 [ ]
MIAT [ , ]
LncRNA RP11-728F11 [ ]
lncRNA RP5-833A20.1 [ ]
lncRNA ANRIL [ ]
LASER [ ]
lncRNA ENST00000416361 [ , ]
LncRNA RAPIA [ , ]
NEAT1 [ , , , , ]
LOC286367 [ ]
HOXC-AS1 [ ]
LncARSR [ , ]
BM450697 [ ]
GAS5 [ , ]
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Gluba-Sagr, A.; Franczyk, B.; Rysz-Górzyńska, A.; Olszewski, R.; Rysz, J. The Role of Selected lncRNAs in Lipid Metabolism and Cardiovascular Disease Risk. Int. J. Mol. Sci. 2024 , 25 , 9244. https://doi.org/10.3390/ijms25179244

Gluba-Sagr A, Franczyk B, Rysz-Górzyńska A, Olszewski R, Rysz J. The Role of Selected lncRNAs in Lipid Metabolism and Cardiovascular Disease Risk. International Journal of Molecular Sciences . 2024; 25(17):9244. https://doi.org/10.3390/ijms25179244

Gluba-Sagr, Anna, Beata Franczyk, Aleksandra Rysz-Górzyńska, Robert Olszewski, and Jacek Rysz. 2024. "The Role of Selected lncRNAs in Lipid Metabolism and Cardiovascular Disease Risk" International Journal of Molecular Sciences 25, no. 17: 9244. https://doi.org/10.3390/ijms25179244

Article Metrics

Further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Lipoprotein lipase and lipolysis: central roles in lipoprotein metabolism and atherogenesis

Affiliation.

  • 1 Department of Medicine, Columbia University College of Physicians and Surgeons, New York, NY 10032.
  • PMID: 8732771

Although it has been known for over 50 years that lipoprotein lipase (LPL) hydrolyzes triglyceride in chylomicrons, during the past half decade there has been a reinterest in the physiologic and pathophysiologic actions of this enzyme. In part, this has coincided with clinical studies implicating increased postprandial lipemia as a risk factor for atherosclerosis development. In addition, the recent creation of genetically altered mice with hypertriglyceridemia has focused the interest of geneticists and physiologists on the pathophysiology of triglyceride metabolism. As reviewed in this article, it is apparent that the lipolysis reaction is only partially understood. Several factors other than LPL are critical modulators of this process, in part, because the reaction requires the lipoproteins to interact with the arterial or capillary wall. Among the factors that affect this are the apolipoprotein composition of the particles, the size of the lipoproteins, and how LPL is displayed along the endothelial luminal surface. Zilversmit's observation that LPL activity is found in greater amounts in atherosclerotic than normal arteries has led to a large number of experiments linking LPL with atherogenesis. In medium and large arteries LPL is found on the luminal endothelial surface and in macrophage-rich areas within the plaque. LPL actions in both of these locations probably have major effects on the biology of the blood vessel. Possible atherogenic actions for this LPL based on in vitro experiments are reviewed.

PubMed Disclaimer

Similar articles

  • [Lipoprotein lipase (LPL) in the pathogenesis of atherosclerosis]. Skoczyńska A. Skoczyńska A. Postepy Hig Med Dosw. 1998;52(4):347-65. Postepy Hig Med Dosw. 1998. PMID: 9780756 Review. Polish.
  • [Lipoprotein lipase-mediated myopathy: implications for lipid metabolism and atherogenesis]. Merkel M, Radner H, Greten H. Merkel M, et al. Med Klin (Munich). 2003 Dec 15;98(12):679-84. doi: 10.1007/s00063-003-1313-6. Med Klin (Munich). 2003. PMID: 14685668 Review. German.
  • Lipoprotein lipase increases lipoprotein binding to the artery wall and increases endothelial layer permeability by formation of lipolysis products. Rutledge JC, Woo MM, Rezai AA, Curtiss LK, Goldberg IJ. Rutledge JC, et al. Circ Res. 1997 Jun;80(6):819-28. doi: 10.1161/01.res.80.6.819. Circ Res. 1997. PMID: 9168784
  • Resistance of chylomicron and VLDL remnants to post-heparin lipolysis in ApoE-deficient mice: the role of apoE in lipoprotein lipase-mediated lipolysis in vivo and in vitro. Zsigmond E, Fuke Y, Li L, Kobayashi K, Chan L. Zsigmond E, et al. J Lipid Res. 1998 Sep;39(9):1852-61. J Lipid Res. 1998. PMID: 9741698
  • Role of macrophage-derived lipoprotein lipase in lipoprotein metabolism and atherosclerosis. Van Eck M, Zimmermann R, Groot PH, Zechner R, Van Berkel TJ. Van Eck M, et al. Arterioscler Thromb Vasc Biol. 2000 Sep;20(9):E53-62. doi: 10.1161/01.atv.20.9.e53. Arterioscler Thromb Vasc Biol. 2000. PMID: 10978269
  • Transcriptomics and gut microbiome analysis of the edible herb Bidens pilosa as a functional feed additive to promote growth and metabolism in tilapia (Oreochromis spp.). Chen CC, Lin CY, Lu HY, Liou CH, Ho YN, Huang CW, Zhang ZF, Kao CH, Yang WC, Gong HY. Chen CC, et al. BMC Genomics. 2024 Aug 13;25(1):785. doi: 10.1186/s12864-024-10674-8. BMC Genomics. 2024. PMID: 39138417 Free PMC article.
  • Long-chain fatty acids - The turning point between 'mild' and 'severe' acute pancreatitis. Liu Q, Gu X, Liu X, Gu Y, Zhang H, Yang J, Huang Z. Liu Q, et al. Heliyon. 2024 May 22;10(11):e31296. doi: 10.1016/j.heliyon.2024.e31296. eCollection 2024 Jun 15. Heliyon. 2024. PMID: 38828311 Free PMC article. Review.
  • Recent advances in the extraction, purification, structural-property correlations, and antiobesity mechanism of traditional Chinese medicine-derived polysaccharides: a review. Zhi N, Chang X, Wang X, Guo J, Chen J, Gui S. Zhi N, et al. Front Nutr. 2024 Jan 17;10:1341583. doi: 10.3389/fnut.2023.1341583. eCollection 2023. Front Nutr. 2024. PMID: 38299183 Free PMC article. Review.
  • The Role of Triglycerides in Atherosclerosis: Recent Pathophysiologic Insights and Therapeutic Implications. Akivis Y, Alkaissi H, McFarlane SI, Bukharovich I. Akivis Y, et al. Curr Cardiol Rev. 2024;20(2):39-49. doi: 10.2174/011573403X272750240109052319. Curr Cardiol Rev. 2024. PMID: 38288833 Review.
  • Correlation of HbA1c Level with Lipid Profile in Type 2 Diabetes Mellitus Patients Visiting a Primary Healthcare Center in Jeddah City, Saudi Arabia: A Retrospective Cross-Sectional Study. Sharahili AY, Mir SA, ALDosari S, Manzar MD, Alshehri B, Al Othaim A, Alghofaili F, Madkhali Y, Albenasy KS, Alotaibi JS. Sharahili AY, et al. Diseases. 2023 Oct 31;11(4):154. doi: 10.3390/diseases11040154. Diseases. 2023. PMID: 37987265 Free PMC article.

Publication types

  • Search in MeSH

Related information

  • Cited in Books

Grants and funding

  • HL 21006/HL/NHLBI NIH HHS/United States
  • R01-45095/PHS HHS/United States

LinkOut - more resources

Full text sources.

  • Elsevier Science

Other Literature Sources

  • The Lens - Patent Citations

Research Materials

  • NCI CPTC Antibody Characterization Program

Miscellaneous

  • NCI CPTAC Assay Portal

full text provider logo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Warning: The NCBI web site requires JavaScript to function. more...

U.S. flag

An official website of the United States government

The .gov means it's official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you're on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • Browse Titles

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

StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan-.

Cover of StatPearls

StatPearls [Internet].

Lipoprotein lipase deficiency.

Suryakumar Balasubramanian ; Pearl Aggarwal ; Saurabh Sharma .

Affiliations

Last Update: July 3, 2023 .

  • Continuing Education Activity

Lipoprotein lipase deficiency is a rare autosomal recessive genetic disorder of lipid metabolism. It is characterized by severe hypertriglyceridemia and chylomicronemia. Early diagnosis and early dietary modification help the prevention of symptoms and medical complications to manifest. This activity reviews the evaluation and treatment of lipoprotein lipase deficiency and highlights the role of the interprofessional team in evaluating and treating patients with this condition.

  • Identify the etiology of lipoprotein lipase enzyme deficiency.
  • Review the appropriate evaluation of lipoprotein lipase enzyme deficiency.
  • Outline the management options available for lipoprotein lipase enzyme deficiency.
  • Describe the importance of collaboration and communication among the interprofessional team to enhance the delivery of care and improve outcomes for patients affected by lipoprotein lipase deficiency.
  • Introduction

Lipoprotein lipase deficiency is a genetic disorder with an autosomal recessive pattern of inheritance. It usually presents in childhood and is characterized by severe hypertriglyceridemia and chylomicronemia. It is the most common form of chylomicronemia and was formerly known as hyperlipoproteinemia type 1a. [1]  Lipoprotein lipase deficiency was first described by Dr. Burger and Dr. Grutz in 1932.

Lipoprotein lipase deficiency occurs due to the presence of the defective gene for lipoprotein lipase that leads to the reduction or complete absence of lipoprotein lipase enzyme activity. Pathogenic deletions, nonsense mutations, and splice-site variants lead to the formation of an abnormal LPL gene product that leads to absent or truncated LPL enzyme with a defective catalytic activity. More than 220 pathological variants, including 70 percent missense mutations, 18 percent nucleotide insertions and deletions, 10 percent nonsense mutations, and a few splice site variants, have been identified. [2] [3]

Because lipoprotein lipase deficiency has an autosomal recessive pattern of inheritance, the risk of two heterozygote parents to have a child affected with lipoprotein lipase deficiency is 25 percent, and the risk of having a heterozygote child is 50 percent, with each pregnancy. Each of the siblings of an affected individual has a 25 percent chance of being homozygous, a 50 percent chance of being heterozygous(carrier), and a 25 percent chance of being unaffected.

  • Epidemiology

Lipoprotein lipase deficiency is a rare disorder. Its prevalence is approximately 1 in 1,000,000 in the general population. Two LPL mutations G118E and P207L cause complete loss of LPL activity in homozygotes and 50% loss in heterozygotes have been reported in Quebec, Canada. Most cases of lipoprotein lipase deficiency are identified in childhood, usually, before ten years of age, and 25 % of the affected patients are identified during the first year of life. However, some individuals may not develop symptoms until adulthood, like women may present for the first time during pregnancy. [1] [4] Males and females are affected equally.

  • Pathophysiology

The lipoprotein lipase gene is mapped to human chromosome 8p22, divided into ten exons and encodes the enzyme lipoprotein lipase, which is expressed in the adipose tissues and muscles. Lipoprotein lipase (LPL), a 475-aminoacid enzyme, is involved in the hydrolysis of triglyceride-rich lipoproteins, mainly chylomicrons and very high-density lipoproteins (VLDL). The catalytic center of the enzyme has three amino acids, namely, Ser132, Asp156, His241.

LPL activity is regulated by several factors, such as hormones, non-esterified fatty acids, and apolipoproteins. Apo AV, apolipoprotein C-II, insulin, acylation stimulating protein increase the LPL activity, while apolipoprotein C-III and TNF-alpha decrease the LPL activity. After LPL is produced in adipose tissues and muscles, which are the two most important sites of production, it is secreted and translocated to the luminal surface of capillary endothelial cells of extrahepatic tissues. The dietary fat absorbed in the intestines is transported in the form of triglycerides by large lipoproteins, known as chylomicrons. Once the chylomicrons are released into the bloodstream, they receive a lipoprotein known as apolipoprotein C-II from high-density lipoproteins.

Apolipoprotein C-II is a cofactor for the lipoprotein lipase enzyme. The lipoprotein lipase recognizes apolipoprotein C-II and gets activated, which results in the breaking down of the chylomicrons and VLDL triglycerides to nonesterified free fatty acids and 2-monoacylglycerol to be stored as triglyceride in adipose tissues or used as an energy source in muscles. [5] [6]  Lipoprotein lipase is also required for maturation of small particles of high-density lipoproteins into larger particles. [7]

Most of the mutations in the lipoprotein lipase gene are located on exons 2, 5, and 6. The well-known mutations include the following

  • Asn291Ser, located on Exon 6, AAT-->AGT. It is the most common mutation in Whites. It is present in 2 to 5 percent of the heterozygote carrier population among the Whites. It is associated with an increased risk of Alzheimer disease. [8]
  • 93T-->G, located on the promotor. The G allele is the most common mutation seen in South African Black ethnicity (76.4 percent). The T allele is seen in Whites (1.7 percent). [9]
  • Asp9Asn, located on Exon 2, GAC-->AAC. It is less common. (1.56 to 4.4 percent of the population). The homozygous mutation causes only a 20 percent decrease in lipoprotein lipase activity. [10]
  • Gly188glu, located on wxon 5, GGG-->GAG. (0.06 percent of the population). [11]  It leads to homozygous chylomicronemia, most common among French Canadians in Quebec. It is known to have the strongest link leading to an increased risk of coronary artery disease. [11]
  • Ala221del, located on exon 5, 916gCT-->CT. It leads to homozygote chylomicronemia. It is the most common mutation in Japanese ethnicity, also known as LPL Arita. [12]
  • Carriers of Asn291Ser or the combined mutation Asp9Asn/T93G are known to have an increased risk of preeclampsia. [13]
  • Ser447sto, located on exon 9, TCA-->TGA. This is a gain of function mutation leading to an increase in LPL activity. It is associated with a 0.8 fold reduced risk of ischemic heart disease, maximum benefit to those using beta-blockers. [14]  Carriers of this mutation have lower plasma triglyceride levels and higher high-density lipoprotein levels. Therefore they have a lower incidence of cardiovascular disease as compared to non-carriers of this mutation. [15]

A reduction or elimination of lipoprotein lipase enzyme activity prevents the break down of triglycerides. Therefore, there is an accumulation of the triglycerides in the blood and tissues, leading to the clinical manifestations of lipoprotein lipase deficiency. [1]  In homozygous individuals, the serum triglyceride levels may reach 10,000 mg/dL or higher. In heterozygous individuals, the serum triglyceride levels may range between 200 to 750 mg/dL.

According to many studies, Lipoprotein lipase deficiency is known to have no atherogenic potential because it causes a low level of low-density lipoproteins. But some studies show that a defect in lipolysis can be a risk factor for premature atherosclerosis. These studies were based on other metabolic disturbances, which have been described as follows. Lipoprotein lipase deficiency also leads to increased serum triglycerides and low high-density lipoproteins. Also, the postprandial clearance of triglycerides is delayed, exposing them to lipoproteins, thus leading to oxidative damage. Even reverse cholesterol transport is impaired as the high-density lipoprotein has structural changes, leading to increased and faster clearance. [16] [7]  

Thus the lipoprotein profile of the patients of lipoprotein lipase deficiency is similar to the postprandial profile in which atherogenic particles are produced, predisposing to atherosclerosis. [17]  However, the association of atherogenesis with lipoprotein lipase deficiency remains debated.

  • History and Physical

Lipoprotein lipase deficiency usually presents with the following:

  • Abdominal pain is the most common presentation and occurs due to acute pancreatitis, intensity varying from mild to incapacitating (mimicking acute abdomen). The attacks of acute pancreatitis are recurrent and usually culminate in chronic pancreatitis. [18]  Although the mechanism of acute pancreatitis following hypertriglyceridemia in LPL enzyme deficiency is not fully understood, oral antioxidants reduce the frequency of pancreatitis episodes, implying the role of oxidant damage in causing acute pancreatitis. [19] [20]  The risk of acute pancreatitis is about 5% when the serum triglycerides reach 1000 mg/dL or higher, and 10% to 20% when the serum triglycerides reach 2000 mg/dL or higher. [21] [22]
  • Xanthomas - In about 50 % of the individuals with lipoprotein lipase deficiency, eruptive xanthomas, which are yellow papules of around 1 mm in size, appear mostly on the trunk, knees, buttocks, and extensor surfaces of arms. These may coalesce to form larger patches. The formation of xanthomas occurs as a result of extravascular phagocytosis of chylomicrons by macrophages and deposition in the skin. They happen when plasma triglyceride levels rise above 2000 mg/dL and disappear when the plasma triglyceride levels return to normal. They are painless unless exposed to repetitive abrasion. [1]
  • Loss of appetite
  • Nausea, vomiting
  • Hepatomegaly and splenomegaly occur as a result of markedly increased plasma triglyceride levels. The excessive chylomicrons in the bloodstream are ingested by macrophages(phagocytes), which then travel to the liver and spleen. The fatty cells accumulate in the liver and spleen, leading to an increase in their size. [1]
  • Retinalis lipemia - the retinal arterioles, venules, and fundus appear pale-pink as a result of scattering of light by the large chylomicrons. This change is reversible, and the vision gets spared. Retinalis lipemia is seen when the plasma triglyceride levels rise above 2500 mg/dL. When the triglyceride levels are between 2500 and 3499 mg/dL, the peripheral vessels of the retina appear thin. At triglyceride levels of 3500 to 5000 mg/dL, the retinal vessels at the posterior pole appear creamy in color. At triglyceride levels above 5000 mg/dL, the retina appears salmon in color with the creamy appearance of the retinal venules and arteries. [23] [24] [1]
  • Neuropsychiatric changes like mild cognitive dysfunction, depression, and memory loss have been reported. These changes are reversible. [25] [5]

Infants may present additionally with the following:

  • Abdominal pain which appears as colic
  • Irritability
  • Intestinal bleeding
  • Failure to thrive [23]

In women, the presentation may be delayed until pregnancy. A woman with lipoprotein lipase deficiency may present with marked signs and symptoms during pregnancy due to increased uptake of triglyceride-rich lipoproteins by the macrophages, owing to the increased apolipoprotein E during pregnancy. [4]

The severity of the clinical presentation of lipoprotein lipase deficiency correlates with the chylomicron levels.

Lipoprotein lipase deficiency is suspected in young individuals with the following clinical findings in addition to the supportive laboratory findings.

Clinical Findings

  • Recurrent attacks of acute pancreatitis
  • Eruptive xanthoma
  • Hepatosplenomegaly
  • Retinalis lipemia

Lab Findings

  • Milky appearing plasma. This occurs due to the impaired clearance of the chylomicrons from the plasma.
  • Plasma triglyceride levels exceeding 2000 mg/dL. These levels are considered regardless of the fasting state.

The diagnosis of lipoprotein lipase deficiency is established by molecular genetic testing, which identifies the proband by identifying the biallelic pathogenic variants in the lipoprotein lipase gene. [26] Two test methods are used, which include:

  • Sequence analysis - This identifies 97 % of probands with the biallelic pathogenic variants.
  • Gene targeted duplication/deletion analysis - this identifies 4% of probands with biallelic pathogenic variants.

Molecular genetic testing can be done for only lipoprotein lipase gene identification, or to identify the other four genes as well, which can lead to chylomicronemia. It is described as follows :

  • Single gene testing - The sequence analysis of only the lipoprotein lipase gene is performed. If one or no pathogenic variant is found, the sequence analysis is followed by the gene-targeted duplication/deletion analysis.
  • Multigene panel - It includes the lipoprotein lipase gene as well as of the other four genes, namely, apolipoprotein C- II gene, apolipoprotein A-V gene, lipase maturation factor 1 gene, and glycosylphosphatidylinositol-anchored high-density lipoprotein binding protein gene, for the chylomicronemia syndrome. Sequence analysis and gene-targeted duplication/deletion, any method may be used. [1] [26]

Measurement of Lipoprotein Lipase Activity

  • Assay of the lipoprotein lipase activity in the plasma, after ten minutes of post Intravenous administration of heparin. Here the post heparin plasma is added to VLDL substrate and the liberated free fatty acids are assayed, and the lipoprotein lipase activity is expressed as micromoles of the released free fatty acids released per minute per liter of the plasma, that is, micromoles/l/min, after subtraction of hepatic lipase activity. Carriers of pathogenic mutations in either the lipoprotein lipase enzyme or its cofactor, that is, apolipoprotein C-II, will lead to undetectable lipoprotein lipase activity in the post heparin plasma which is a diagnostic of familial lipoprotein lipase deficiency. [27] [28] [26] [28]
  • Biopsies of the adipose tissue may also be used to determine the lipoprotein lipase activity directly.
  • Treatment / Management

Medical nutrition therapy involves following a fat-restricted diet to keep the individual diagnosed with lipoprotein lipase deficiency free of signs and symptoms. The targeted goal is to keep the plasma triglyceride levels below 2000 mg/dL, with the greatest benefit obtained when the plasma triglyceride levels are kept below 1000 mg/dL. This can be achieved by restricting the dietary fat intake to not above 20 g/day or 15% of total energy intake.

Fish oil supplements are not beneficial and are contraindicated in lipoprotein lipase deficiency, unlike the disorders of excess hepatic triglyceride production. This is because fish oils contribute to chylomicrons. Certain agents like alcohol, oral estrogens, beta-adrenergic blockers, diuretics, selective serotonin reuptake inhibitors, isotretinoin, are avoided as they are known to increase the endogenous triglyceride levels. [1] [26]  For individuals who take a very low-fat diet, it is recommended that they should supplement fat-soluble vitamins, that is, vitamin A, D, E, K, and minerals in their diet.

Acute pancreatitis associated with lipoprotein lipase deficiency is treated in the same manner as acute pancreatitis due to other causes. Prevention of recurrent acute pancreatitis helps to decrease the risk of secondary complications of pancreatitis like diabetes mellitus.

Plasma triglycerides are followed overtime of the affected individual as this helps to evaluate the success of the fat-restricted diet. 

Pregnant women need to follow an extreme fat-restricted diet, that is, less than 2 g/day, especially during the second and third trimesters. This, along which close monitoring of the plasma triglyceride levels, leads to the delivery of normal infants with normal plasma levels of essential fatty acids. A combination of a very low-fat diet with the use of gemfibrozil has been safely implicated in pregnancy. [1] [29]

For family planning, it is appropriate to offer genetic counseling to young adults who are affected, are carriers, or at risk of being a carrier.

Lipoprotein lipase (LPL) gene therapy, that is, alipogene tiparvovec gene therapy, consists of the LPL Ser447X variant in a genetically engineered adeno-associated virus genotype 1 (alipogene tiparvovec). The intramuscular administration of the adenovirus vector introduces a functional copy of the lipoprotein lipase gene into the patient's muscle cells, thus lowering the fasting triglyceride levels. [30] [31] [32]  The maximum benefit is for individuals with the highest risk of complications. However, due to low demand from the patient community, it has been taken off the market. [1] [33] [34]

Currently, there are some promising treatment approaches, that include the following:

  • Pradigastat is a diacylglycerol  O -acyltransferase 1 (DGAT1) inhibitor that is orally administered. It functions by catalyzing the final step in triglyceride synthesis and decreases the chylomicron-triglyceride secretion. [23] [35] [23]
  • Evinacumab is a monoclonal antibody to angiopoietin-like protein 3 (ANGPTL3). It is an inhibitor of LPL and endothelial lipase. [23]
  • Differential Diagnosis

LPL deficiency is considered in young individuals with chylomicronemia and triglyceride levels of more than 2000 mg/dl. However, it has been found that such individuals do not necessarily have familial LPL deficiency. Instead, they may have one of the more common genetic disorders of triglyceride metabolism, like familial combined hyperlipidemia and monogenic familial hypertriglyceridemia, or there may be some secondary causes leading to hypertriglyceridemia. [1]

Other than LPL deficiency, that constitutes 95.0 percent of primary monogenic variants of chylomicronemia, the differential diagnoses for primary monogenic chylomicronemia include:

  • Familial apolipoprotein C- II deficiency - severe chylomicronemia in childhood or adolescence, constitutes 2.0 percent of monogenic variants.
  • Familial apolipoprotein A-V deficiency-chylomicronemia in late adulthood constitutes 0.6 percent of monogenic variants.
  • Familial lipase maturation factor 1 deficiency- chylomicronemia in late adulthood, constitutes 0.4 percent of monogenic variants.
  • Familial glycosylphosphatidylinositol-anchored high-density lipoprotein binding protein 1 deficiency-Chylomicronemia in late adulthood, constitutes 2.0 percent of monogenic variants. [1] [36] [37] [38]  

The following secondary causes can also cause hypertriglyceridemia:

  • Diabetes mellitus
  • Paraproteinemia and lymphoproliferative disorders
  • Alcohol use
  • Estrogen therapy
  • Drugs like selective serotonin receptor reuptake inhibitors, glucocorticoids, atypical antipsychotics, isotretinoin, certain antihypertensives [1]

Medical nutrition therapy is the mainstay for the treatment of lipoprotein lipase deficiency. Thus, treatment success depends on the acceptance of the fat-restricted diet of the individual affected. The ultimate prognosis of lipoprotein lipase deficiency appears to be good with high compliance for a fat-restricted diet that leads to a decrease in plasma triglyceride levels. The enlarged liver and spleen usually return to normal size within one week of lowering down of the plasma triglyceride levels, eruptive xanthomas clear within a few weeks to months. [1]  In lipoprotein lipase deficiency, even with a history of recurrent attacks of acute pancreatitis, the pancreatic function declines slowly, so it is not associated with high mortality. [20] [1]

  • Complications

In an individual with lipoprotein lipase deficiency, recurrent attacks of acute pancreatitis lead to chronic pancreatitis. The secondary complications of chronic pancreatitis are as follows:

  • Steatorrhea, that is, stools containing excess fat due to fat malabsorption
  • Pancreatic calcifications

However, these complications are rare in an individual with lipoprotein lipase deficiency. Even if these complications occur, they are rare before mid-age. Pancreatitis might be rarely associated with serious complications like total pancreatic necrosis and death. [1]

  • Deterrence and Patient Education

The quality of life of the individuals affected with lipoprotein lipase deficiency is poor, mainly due to recurrent attacks of acute pancreatitis. Patients and their family members are noted to be anxious, depressed, and frustrated during and after hospitalizations for the attacks of pancreatitis. Recurrent hospitalizations affect various aspects of daily life, for example, work-life due to absenteeism, financial implications, and increased dependency on family and friends for support. [39] [40]

It is essential to educate the patient about the importance of following a strict fat-restricted diet to get relief from the signs and symptoms of lipoprotein lipase deficiency and to prevent its secondary manifestations. Daily life modifications, for example, using sources of medium-chain fatty acids for cooking, which get absorbed into the portal vein directly without getting incorporated into the chylomicron triglyceride, should be encouraged. A dietician consult could be helpful to achieve the goal of required daily fat consumption in the diet. Periodic follow-up for diet review along with the plasma triglyceride levels can help ensure therapy success. [1]

Genetic counseling, that is, educating the affected individual about the nature, mode of inheritance, and the impact of this condition, is also quite important. This is because early diagnosis of this condition and early dietary modification helps the prevention of symptoms and medical complications to manifest. [37]  This will help the patient to make informed medical and personal decisions.

  • Enhancing Healthcare Team Outcomes

Though lipoprotein lipase deficiency is a rare genetic disorder, its implications on the life of an affected individual are quite debilitating. Most of it is attributed to the recurrent attacks of acute pancreatitis that leads to multiple hospitalizations. The manifestations of this disorder, the need to follow a strict fat-restricted diet and associated impaired psychosocial functioning have a poor impact on the Health-Related Quality of Life (HRQoL). The lack of proven effective and cost-effective therapies further increase the disease burden. 

 The unmet need for proper and consistent dietary advice, as well as the education of the patients and their friends and family, can be addressed by the following measures:

  • Healthcare professionals must obtain complete knowledge and understanding of the disorder. This will ensure appropriate diets to be followed during hospitalizations and allow patients to better understand their condition and follow a strict diet at home.
  • Emotional support services and support groups that would help to decrease the uncertainty and the fear of having future attacks of acute pancreatitis. 
  • Provision of materials offering the sources of information on the disorder. It will ease the anxiety and tension that the patients and their friends and family face.

It is essential to consult with an interprofessional team of specialists that include a pediatrician, gastroenterologist, surgeon, endocrinologist, ophthalmologist, gynecologist, and a dietician. The nurses are also a vital member of the interprofessional group as they will monitor the patient's vital signs and assist with the education of the patient and family. The pharmacist will ensure that the patient is on the right analgesics in the event of attacks of acute pancreatitis. The radiologist also plays a vital role in determining the cause of abdominal pain. In cases where evidence is not definitive or minimal, expert opinion from the specialist may be utilized to recommend the type of diagnostic test or treatment. However, to improve outcomes, prompt consultation with an interprofessional group of specialists is recommended.

  • Review Questions
  • Access free multiple choice questions on this topic.
  • Comment on this article.

Disclosure: Suryakumar Balasubramanian declares no relevant financial relationships with ineligible companies.

Disclosure: Pearl Aggarwal declares no relevant financial relationships with ineligible companies.

Disclosure: Saurabh Sharma declares no relevant financial relationships with ineligible companies.

This book is distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International (CC BY-NC-ND 4.0) ( http://creativecommons.org/licenses/by-nc-nd/4.0/ ), which permits others to distribute the work, provided that the article is not altered or used commercially. You are not required to obtain permission to distribute this article, provided that you credit the author and journal.

  • Cite this Page Balasubramanian S, Aggarwal P, Sharma S. Lipoprotein Lipase Deficiency. [Updated 2023 Jul 3]. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan-.

In this Page

Bulk download.

  • Bulk download StatPearls data from FTP

Related information

  • PMC PubMed Central citations
  • PubMed Links to PubMed

Similar articles in PubMed

  • Autoantibodies against GPIHBP1 as a Cause of Hypertriglyceridemia. [N Engl J Med. 2017] Autoantibodies against GPIHBP1 as a Cause of Hypertriglyceridemia. Beigneux AP, Miyashita K, Ploug M, Blom DJ, Ai M, Linton MF, Khovidhunkit W, Dufour R, Garg A, McMahon MA, et al. N Engl J Med. 2017 Apr 27; 376(17):1647-1658. Epub 2017 Apr 5.
  • Review The role of registries in rare genetic lipid disorders: Review and introduction of the first global registry in lipoprotein lipase deficiency. [Atherosclerosis. 2017] Review The role of registries in rare genetic lipid disorders: Review and introduction of the first global registry in lipoprotein lipase deficiency. Steinhagen-Thiessen E, Stroes E, Soran H, Johnson C, Moulin P, Iotti G, Zibellini M, Ossenkoppele B, Dippel M, Averna MR, et al. Atherosclerosis. 2017 Jul; 262:146-153. Epub 2016 Aug 21.
  • Premature atherosclerosis in patients with familial chylomicronemia caused by mutations in the lipoprotein lipase gene. [N Engl J Med. 1996] Premature atherosclerosis in patients with familial chylomicronemia caused by mutations in the lipoprotein lipase gene. Benlian P, De Gennes JL, Foubert L, Zhang H, Gagné SE, Hayden M. N Engl J Med. 1996 Sep 19; 335(12):848-54.
  • Review Familial Lipoprotein Lipase Deficiency. [GeneReviews(®). 1993] Review Familial Lipoprotein Lipase Deficiency. Burnett JR, Hooper AJ, Hegele RA. GeneReviews(®). 1993
  • An incomplete form of familial lipoprotein lipase deficiency presenting with type I hyperlipoproteinemia. [Am J Clin Pathol. 1987] An incomplete form of familial lipoprotein lipase deficiency presenting with type I hyperlipoproteinemia. Berger GM. Am J Clin Pathol. 1987 Sep; 88(3):369-73.

Recent Activity

  • Lipoprotein Lipase Deficiency - StatPearls Lipoprotein Lipase Deficiency - StatPearls

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

Connect with NLM

National Library of Medicine 8600 Rockville Pike Bethesda, MD 20894

Web Policies FOIA HHS Vulnerability Disclosure

Help Accessibility Careers

statistics

IMAGES

  1. Figure 2 from Lipoprotein Lipase: A General Review

    lipoprotein lipase experiment

  2. Lipoprotein lipase catalyzes the hydrolysis of triglycerides

    lipoprotein lipase experiment

  3. Lipoprotein lipase activity in milk stored at 4°C or 31°C for 4 to 24 h

    lipoprotein lipase experiment

  4. Models for the attachment of lipoprotein lipase to the cell surface

    lipoprotein lipase experiment

  5. Localization of lipoprotein lipase (LPL),...

    lipoprotein lipase experiment

  6. Roar Lipoprotein Lipase Activity Kit

    lipoprotein lipase experiment

COMMENTS

  1. Lipoprotein Lipase and Its Regulators: An Unfolding Story

    Lipoprotein lipase (LPL): A Central Player in Systemic Lipid Metabolism. LPL was first discovered in 1943 by Dr. Paul Hanh in dogs as "heparin-activated clearing factor" [] and renamed to "lipoprotein lipase" by Dr. Edward Korn in 1955 [2, 3].It was later revealed that LPL is a highly conserved protein among mammalian species with a molecular mass of 50 kDa [].

  2. Biochemistry, Lipoprotein Lipase

    Lipoprotein lipase (LPL) is an extracellular enzyme on the vascular endothelial surface that degrades circulating triglycerides in the bloodstream. These triglycerides are embedded in very low-density lipoproteins (VLDL) and chylomicrons traveling through the bloodstream. The role of lipoprotein lipase is significant in understanding the pathophysiology of type one familial dyslipidemias, or ...

  3. Lipoprotein lipase

    Lipoprotein lipase (LPL) (EC 3.1.1.34, systematic name triacylglycerol acylhydrolase (lipoprotein-dependent)) is a member of the lipase gene family, which includes pancreatic lipase, hepatic lipase, and endothelial lipase.It is a water-soluble enzyme that hydrolyzes triglycerides in lipoproteins, such as those found in chylomicrons and very low-density lipoproteins (VLDL), into two free fatty ...

  4. Lipoprotein Lipase and Its Regulators: An Unfolding Story

    Lipoprotein lipase (LPL) is one of the most important factors in systemic lipid partitioning and metabolism. It mediates intravascular hydrolysis of triglycerides packed in lipoproteins such as chylomicrons and very-low-density lipoprotein (VLDL). Since its initial discovery in the 1940s, its biology and pathophysiological significance have been well characterized. Nonetheless, several studies ...

  5. Lipoprotein lipase: structure, function, regulation, and role in

    Abstract. Lipoprotein lipase (LPL) catalyses the hydrolysis of the triacylglycerol component of circulating chylomicrons and very low density lipoproteins, thereby providing non-esterified fatty acids and 2-monoacylglycerol for tissue utilisation. Research carried out over the past two decades have not only established a central role for LPL in ...

  6. Lipoprotein lipase

    Lipoprotein lipase (LPL) regulates the plasma levels of triglyceride and HDL. Three aspects are reviewed. 1) Clinical implications of human LPL gene variations: common mutations and their effects on plasma lipids and coronary heart disease are discussed. 2) LPL actions in the nervous system, liver, and heart: the discussion focuses on LPL and tissue lipid uptake.3) LPL gene regulation: the LPL ...

  7. PDF Lipoprotein Lipase and Its Regulators: An Unfolding Story

    Lipoprotein lipase (LPL) is one of the most important factors in systemic lipid partitioning and metabolism. It mediates intravascular hydrolysis of triglycerides packed in lipoproteins such as chylomicrons and very-low-density lipoprotein (VLDL). Since its initial discovery in the 1940s, its biology and pathophysiological significance have ...

  8. PDF A unified model for regulating lipoprotein lipase activity

    The regulation of triglyceride (TG) tissue distribution, storage, and utilization, a fundamental process of energy homeostasis, critically depends on lipoprotein lipase (LPL). We review the intricate mechanisms by which LPL activity is regu-lated by angiopoietin-like proteins (ANGPTL3, 4, 8), apolipoproteins (APOA5, APOC3, APOC2), and the cAMP ...

  9. Regulation of lipoprotein lipase-mediated lipolysis of triglycerides

    Lipoprotein lipase (LpL)-mediated lipolysis has long been known to be the rate-limiting step in the removal of triglyceride from the blood stream. Within the past decade, LpL and its regulators have garnered increased attention in part because genes that modulate LpL activity correlate with the risk of cardiovascular disease (CVD) events.

  10. A unified model for regulating lipoprotein lipase activity

    Abstract. The regulation of triglyceride (TG) tissue distribution, storage, and utilization, a fundamental process of energy homeostasis, critically depends on lipoprotein lipase (LPL). We review the intricate mechanisms by which LPL activity is regulated by angiopoietin-like proteins (ANGPTL3, 4, 8), apolipoproteins (APOA5, APOC3, APOC2), and ...

  11. Lipoprotein lipase: from gene to obesity

    lipoprotein lipase (LPL) plays a major role in the metabolism and transport of lipids. It is the enzyme responsible for the hydrolysis of core triglycerides (TGs) in chylomicrons and very low-density lipoproteins (VLDLs), producing chylomicron remnants and intermediate-density lipoproteins (IDLs), respectively (54, 86).Besides its hydrolytic activity, LPL can interact with lipoproteins to ...

  12. The Regulation of Triacylglycerol Metabolism and Lipoprotein Lipase

    The Regulation of Triacylglycerol Metabolism and Lipoprotein Lipase Activity. Yi Wen, Yi Wen. Lilly Research Laboratories, Eli Lilly and Company, Indianapolis, IN, 46285 USA ... to maintain a pool of TG within circulating lipoproteins that can be hydrolyzed in a tissue-specific manner by lipoprotein lipase (LPL) to enable the delivery of fatty ...

  13. Lipoprotein Lipase and Its Regulators: An Unfolding Story

    Abstract. Lipoprotein lipase (LPL) is one of the most important factors in systemic lipid partitioning and metabolism. It mediates intravascular hydrolysis of triglycerides packed in lipoproteins such as chylomicrons and very-low-density lipoprotein (VLDL). Since its initial discovery in the 1940s, its biology and pathophysiological ...

  14. Lipoprotein lipase transport in plasma: role of muscle and adipose

    Lipoprotein lipase (LPL) is synthesized in tissues involved in fatty acid metabolism such as muscle and adipose tissue. LPL is also found in the circulation, but is mostly lipolytically inactive. The proportion of active circulating LPL increases after a fatty meal. We investigated the release of active and inactive LPL from adipose tissue and muscle in the fasting and postprandial states.

  15. The Effects of Weight Loss on the Activity and Expression of Adipose

    When adipose lipoprotein lipase has been measured in obese subjects after they have fasted, an increase in lipoprotein lipase activity per adipocyte has been demonstrated consistently. 5-11 ...

  16. The Importance of Lipoprotein Lipase Regulation in Atherosclerosis

    Abstract. Lipoprotein lipase (LPL) plays a major role in the lipid homeostasis mainly by mediating the intravascular lipolysis of triglyceride rich lipoproteins. Impaired LPL activity leads to the accumulation of chylomicrons and very low-density lipoproteins (VLDL) in plasma, resulting in hypertriglyceridemia.

  17. Determination of lipoprotein lipase activity using a novel fluorescent

    A novel, real-time, homogeneous fluorogenic lipoprotein lipase (LPL) assay was developed using a commercially available substrate, the EnzChek lipase substrate, which is solubilized in Zwittergent. The triglyceride analog substrate does not fluoresce, owing to apposition of fluorescent and fluorescent quenching groups at the sn-1 and sn-2 positions, respectively, fluorescence becoming ...

  18. Lipoprotein Lipase and Its Regulators: An Unfolding Story

    Lipoprotein lipase (LPL) is one of the most important factors in systemic lipid partitioning and metabolism. It mediates intravascular hydrolysis of triglycerides packed in lipoproteins such as chylomicrons and very-low-density lipoprotein (VLDL). Since its initial discovery in the 1940s, its biology and pathophysiological significance have ...

  19. GPIHBP1 and Lipoprotein Lipase, Partners in Plasma Triglyceride

    Lipoprotein lipase (LPL), identified in the 1950s, has been studied intensively by biochemists, physiologists, and clinical investigators. These efforts uncovered a central role for LPL in plasma triglyceride metabolism and identified LPL mutations as a cause of hypertriglyceridemia. By the 1990s, with an outline for plasma triglyceride metabolism established, interest in triglyceride ...

  20. Determination of lipoprotein lipase activity using a novel fluorescent

    A novel, real-time, homogeneous fluorogenic lipoprotein lipase (LPL) assay was developed using a commercially available substrate, the EnzChek lipase substrate, which is solubilized in Zwittergent. The triglyceride analog substrate does not fluoresce, owing to apposition of fluorescent and fluorescent quenching groups at the sn-1 and sn-2 ...

  21. Lipoprotein Lipase: Is It a Magic Target for the Treatment of

    Lipoprotein lipase (LPL) is a key regulator for TGs that hydrolyzes TGs to glycerol and free fatty acids in lipoprotein particles for lipid storage and consumption in peripheral organs. A deeper understanding of human genetics has enabled the identification of proteins regulating the LPL activity, which include the apolipoproteins and ...

  22. A negatively charged cluster in the disordered acidic domain ...

    GPIHBP1 is a membrane protein of endothelial cells that transports lipoprotein lipase (LPL), the key enzyme in plasma triglyceride metabolism, from the interstitial space to its site of action on ...

  23. IJMS

    Lipid disorders increase the risk for the development of cardiometabolic disorders, including type 2 diabetes, atherosclerosis, and cardiovascular disease. Lipids levels, apart from diet, smoking, obesity, alcohol consumption, and lack of exercise, are also influenced by genetic factors. Recent studies suggested the role of long noncoding RNAs (lncRNAs) in the regulation of lipid formation and ...

  24. Subclavian vein

    The subclavian vein is a paired large vein, one on either side of the body, that is responsible for draining blood from the upper extremities, allowing this blood to return to the heart.The left subclavian vein plays a key role in the absorption of lipids, by allowing products that have been carried by lymph in the thoracic duct to enter the bloodstream.

  25. Lipoprotein lipase and lipolysis: central roles in lipoprotein

    Abstract. Although it has been known for over 50 years that lipoprotein lipase (LPL) hydrolyzes triglyceride in chylomicrons, during the past half decade there has been a reinterest in the physiologic and pathophysiologic actions of this enzyme. In part, this has coincided with clinical studies implicating increased postprandial lipemia as a ...

  26. Lipoprotein Lipase Deficiency

    Lipoprotein lipase deficiency is a genetic disorder with an autosomal recessive pattern of inheritance. It usually presents in childhood and is characterized by severe hypertriglyceridemia and chylomicronemia. It is the most common form of chylomicronemia and was formerly known as hyperlipoproteinemia type 1a.[1] Lipoprotein lipase deficiency was first described by Dr. Burger and Dr. Grutz in ...