IMR Press / FBL / Volume 28 / Issue 1 / DOI: 10.31083/j.fbl2801010
Open Access Review
Cellular Uptake, Metabolism and Sensing of Long-Chain Fatty Acids
Show Less
1 China International Science and Technology Cooperation Base of Food Nutrition/Safety and Medicinal Chemistry, College of Bioengineering, Tianjin University of Science & Technology, 300457 Tianjin, China
2 Inner Mongolia Academy of Agricultural & Animal Husbandry Science, 010031 Hohhot, Inner Mongolia, China
3 School of life Sciences, Inner Mongolia University, 010070 Hohhot, Inner Mongolia, China
*Correspondence: yupeng@tust.edu.cn (Peng Yu)
Academic Editor: Ananda Ayyappan Jaguva Vasudevan
Front. Biosci. (Landmark Ed) 2023, 28(1), 10; https://doi.org/10.31083/j.fbl2801010
Submitted: 3 October 2022 | Revised: 8 November 2022 | Accepted: 14 November 2022 | Published: 16 January 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Fatty acids (FAs) are critical nutrients that regulate an organism’s health and development in mammal. Long-chain fatty acids (LCFAs) can be divided into saturated and unsaturated fatty acids, depending on whether the carbon chain contains at least 1 double bond. The fatty acids that are required for humans and animals are obtained primarily from dietary sources, and LCFAs are absorbed from outside of cells in mammals. LCFAs enter cells through several mechanisms, including passive diffusion and protein-mediated translocation across the plasma membrane, the latter in which FA translocase (FAT/CD36), plasma membrane FA-binding protein (FABPpm), FA transport protein (FATP), and caveolin-1 are believed to have important functions. The LCFAs that are taken up by cells bind to FA-binding proteins (FABPs) and are transported to the specific organelles, where they are activated into acyl-CoA to target specific metabolic pathways. LCFA-CoAs can be esterified to phospholipids, triacylglycerol, cholesteryl ester, and other specialized lipids. Non-esterified free fatty acids are preferentially stored as triacylglycerol molecules. The main pathway by which fatty acids are catabolized is β-oxidation, which occurs in mitochondria and peroxisomes. stearoyl-CoA desaturase (SCD)-dependent and Fatty acid desaturases (FADS)-dependent fatty acid desaturation pathways coexist in cells and provide metabolic plasticity. The process of fatty acid elongation occurs by cycling through condensation, reduction, dehydration, and reduction. Extracellular LCFA can be mediated by membrane protein G protein-coupled receptor 40 (GPR40) or G protein-coupled receptor 120 (GPR120) to activate mammalian target of rapamycin complex 1 (mTORC1) signaling, and intracellular LCFA’s sensor remains to be determined. The crystal structures of a phosphatidic acid phosphatase and a membrane-bound fatty acid elongase-condensing enzyme and other LCFA-related proteins provide important insights into the mechanism of utilization, increasing our understanding of the cellular uptake, metabolism and sensing of LCFAs.

Keywords
fatty acid sensing
lipid metabolism
mTORC1
homeostasis
1. Introduction

In addition to serving as building blocks for lipid synthesis, fatty acids (FAs) are needed for membrane function, energy storage, and signaling. An adequate supplies of fatty acids is important for maintaining metabolism and energy homeostasis in cells. Animal cells obtain fatty acids primarily by extracellular uptake, de novo synthesis, and hydrolytic cleavage of ester bonds in triacylglycerol stored in tissues [1, 2, 3]. Moreover, essential fatty acids can be synthesized by rumen microorganisms in ruminant animals [4, 5, 6].

Fatty acids are classified by their carbon (C) chain length and degree of desaturation, each of which differs content in living cells or milk [7, 8]. Long-chain (LC) fatty acids refer to fatty acids with a chain length of 11/12–20 carbons and are precursors of various lipids that participate in various physiological processes, e.g., cellular metabolism, energy homeostasis, and cell proliferation [9, 10]. In general, Long-chain fatty acids (LCFAs) synthesis was found to vary, depending on the tissue type, the contribution of fatty acid synthesis de novo to the whole fatty acid pool is not dominant, and they are absorbed primarily from the outside of cells [11, 12]. The LCFAs that are taken up by cells are managed in many metabolic cascades, including their release from the inner leaflet of the plasma membrane, transport to specific organelles, and activation in cells. Mammalian cells possess the ability to properly sense both extracellular and intracellular nutrients for the maintenance of metabolic homeostasis, including lipid homoeostasis. However, the sensing mechanism by which fatty acids are taken up and used by human or animal cells is not fully understood.

2. Fatty Acid Classification and Dietary Sources

Fatty acids are components of intracellular lipids, which store energy in the form of triacylglycerol in mammalian cells. Fatty acids are classified by their carbon (C) chain length. Short-chain (SC) FAs have a chain length of between 1 and 4 C atoms, comprising acetic (2:0), propionic (3:0), and butyric (4:0) acids, and medium-chain (MC) FAs have lengths of 6-10 Cs, including caproic (6:0), caprylic (8:0), and capric (10:0) acids [13]. LCFAs have chain length of 11/12–20 Cs, e.g., lauric (12:0), myristic (14:0), palmtic (16: 0), stearic (18:0), and arachidic acid (20:0), of which C16/18 LCFAs are the most abundant FA species in mammalian cells [13, 14]. Very long-chain fatty acids (VLCFAs) are defined as FAs with C 22, C 24, and C 26, such as behenic (22:0), lignoceric (24:0), cerotic (26:0) and are less abundant than LCFAs. C22 and C24 VLCFAs are ubiquitous in the body, and C 26 VLCFAs are usually subclassified into ultra (U)-LCFAs, which are tissue-specific and found in the skin, retina, meibomian gland, testis, and brain [15]. VLCFAs can be categorized into saturated (SFAs), monounsaturated (MUFAs), and polyunsaturated fatty acids (PUFAs) each of which has specific characteristics [16]. Cis/trans-fatty acids are unsaturated FAs that contain at least 1 double bond in cis or trans, and neither linoleic acid (LA, 18:2n-6) nor alpha (α)-linolenic acid (ALA, 18:3n-3), designated as essential fatty acids (EFAs), cannot be synthesized completely by mammalian cells and thus their consumption is required from dietary sources in order to meet biological demands [17].

The free fatty acids that are required for humans and animals are derived from endogenous synthesis, or come from exogenous sources. Long- and medium-chain fatty acids derived mainly from dietary triacylglycerol, and short-chain fatty acids (SCFAs), also called volatile FAs, produced by gut microbial fermentation—in particular, acetate, propionate and butyrate [18, 19]. Medium-chain fatty acids (MCFAs) are absorbed from dietary plant oils and milk directly into the portal blood, e.g., capric acid in coconut oil [13, 20]. In addition to being provided by the diet, SCFAs and MCFAs can be formed in mammalian and human tissues-primarily the liver, mammary gland, and adipose tissue [20]. LCFAs/LCPUFAs and VLCFAs, including SFAs, MUFAs, and PUFAs, are obtained mainly from the diet [13, 21], whereas sources of unsaturated fatty acids are vegetable oils, e.g., oleic acids in olive oil (Table 1, Ref. [13, 15, 19, 20, 21]) [13].

Table 1.Fatty acids classification, chain length, dietary sources and main tissues for storage.
Fatty acids (FAs) classification Dietary sources In mammalian and human tissues Reference
SCFAs: 1 and 4 C atoms Produced chiefly by anaerobic fermentation and the metabolism of dietary fiber by gut rumen microbes. Primarily the liver, mammary gland, and adipose tissue. [19]
MCFAs: 6-10 C atoms Absorbed from dietary plant oils and milk directly into the portal blood, e.g., coconut oil. Primarily the liver, mammary gland, and adipose tissue. [20]
LCFAs: 11/12-20 C atoms Diet, e.g., olive oil. [13, 21]
VLCFAs: C 22, C 24, C 26 C atoms Obtained mainly from the diet and through elongated and desaturated of endogenous FAs, e.g., peanut oil. Skin, retina, meibomian gland, testis, and brain. [15]
FAs, Fatty Acids; LCFAs, Long-chain fatty acids; MCFAs, Medium-chain fatty acids; SCFAs, Short-chain fatty acids; VLCFAs, Very long-chain fatty acids.

VLCFAs are synthesized through elongation and desaturation of endogenous FAs, such as palmitic acid (C16:0) and stearic acid (C18:0), in animals and humans [15]. In the udder of dairy cows, FAs with C4:0-C14:0 and approximately 50% of C16:0 are synthesized de novo from acetate and β-hydroxybutyrate, and the remaining C16:0 and other longer-chain FAs are absorbed from the diet [22]. Further, C16:0 can be elongated into long-chain FAs that contain 18 carbon atoms or VLCFAs [14].

In addition to external uptake and de novo synthesis, mammalian cells can obtain fatty acids through the hydrolytic cleavage of ester bonds in triacylglycerol from fat stores to maintain fatty acids homeostasis under fasting conditions [23, 24]. Dietary fat was broken down by lipases in small intestine to degrade triacylglycerol (TG) in human. The breakdown of TG generates fatty acids, glycerol and monoglycerides. In the process, three major lipases have been identified, including adipose triglyceride lipase (ATGL), hormonesensitive lipase (HSL) and monoglyceride lipase (MGL), which sequentially performs the TG hydrolysis generating diglycerides (DGs) and FAs, DGs hydrolysis generating monoglycerides (MGs) and MGs hydrolysis generates glycerol and the third FA [23]. In addition, the hydrolysis of cholesteryl esters also generates free fatty acids in human cells. Cholesterol can be derived from dietary sources in the intestine, and do novo synthesized in liver. Both endogenously synthesized and exogenously acquired cholesterol are processed into low-density lipoprotein chiolesterol (LDL-C) in blodstream, which can be taken up by peripheral cells. Excess cholesterol is esterified by acyl coenzyme A: cholesterol acyltransferase (ACAT) to cholesteryl esters for storage in lipid droplets in cells [24]. The cycle of esterification and hydrolysis of cholesterol esters is one of the important element for the lipid homoeostasis in cells. In spite of this, de novo synthesis of fatty acids constitutes a minor but necessary source (Fig. 1).

Fig. 1.

The source of intracellular FAs. In mammalian cells, cellular FAs are generally absorbed from outside of cells; de novo synthesis of FAs is a minor but necessary pathway, and the catabolism of triacylglycerol to complement FAs occurs during nutrient deprivation. FAs, Fatty Acids; ER, endoplasmic reticulum.

3. Cellular Uptake of LCFAs

LCFA uptake by cells is important in maintaining lipid homeostasis. Most LCFAs that circulate in body fluid exist in the form of free LCFA, complexes with albumin, and local lipoprotein lipase (LPL)-mediated LCFA release from membrane bound lipoprotein [25]. The cellular uptake of exogenous LCFAs for use in cells occurs through a cascade, comprising the dissociation of FAs from albumin-FA complexes and their binding to plasma membrane proteins, FA translocation across the plasma membrane, binding cytoplasmic FABP on the inner plasma membrane, and the activation of LCFAs into acyl-CoA, which is necessary for such metabolic processes as TG synthesis and oxidation [17].

The first step is the release of FAs from albumin-FA complexes for presentation to the cell surface [25]—a process that remains poorly understood. The dissociation of FAs from albumin-FA complexes is believed to be facilitated by membrane-associated proteins, including FA translocase (FAT/CD36), plasma membrane FA-binding protein (FABPpm), and FA transport protein (FATP) [26]. FAT/CD36, FABPpm, FATP, and caveolin-1 are thought to mediate the translocation of LCFAs across the plasma membrane.

3.1 FAT/CD36 Appears to be the Most Important Translocator of LCFAs

In 1993, a cDNA clone from a rat adipocyte cDNA library was isolated by screening and implicated in the transport of LCFAs. It was found to be homologous to human CD36 and was termed FAT/CD36 [27]. CD36 is a multifunctional membrane protein, and its relative molecular weight depends on its post-translational modification [28]. Glycosylated CD36 increases the uptake of LCFAs [29]. Its structure is divided into 5 regions: carboxy-terminal (COOH-terminal) and amino-terminal cytoplasmic domains (NH2-terminal), 2 transmembrane regions, and an extracellular domain. The COOH-terminus contains 2 palmitoylation sites and 2 ubiquitination sites, and the NH2-terminus contains only 2 palmitoylation sites—palmitoylated CD36 is located in the lipid rafts of the cell membrane where it mediates adsorption and transport of fatty acids [29, 30].

The extracellular domain is a large, highly glycosylated hydrophobic neck ring and contains 3 pairs of disulfide bonds, 10 glycosylation sites, and 2 phosphorylation sites. These modified sites can interact with a variety of extracellular substances, such as oxidized low-density lipoprotein (ox-LDL) and LCFAs [30, 31]. In dairy cows and goats, CD36 is expressed by the mammary glands and respond to LCFAs to improve milk lipid synthesis [32, 33]. Palmitic acid upregulates CD36 and promotes its translocation from the cytoplasm to plasma membrane in mouse podocytes [34]. CD36 appears to be the most important translocator of FAs, based on current evidence.

3.2 FABPpm Binds to LCFAs with Its Hydrophobic Tail and Facilitates Dissociation from Albumin to Mediate the Translocation of Them

In 1985, FABPpm was isolated and identified by the Berk group from rat liver plasma membranes and jejunal microvillous membranes [35], adipocytes [36], and cardiac myocytes [37]. Subsequently, the protein was determined to be identical to the mitochondrial isoenzyme glutamic-oxaloacetic transaminase (mGOT)/mitochondrial aspartate aminotransferase (mAspAT) [35]. FABPpm is anchored to the outer leaflet of the plasma membrane, with its hydrophobic tail binding to FAs with high affinity, their facilitating dissociation from albumin [17]. The overexpression of FABPpm in vitro or in vivo increases the rate of LCFA transport and metabolism [38, 39, 40]. The mechanism by which FABPpm transports LCFAs is not fully understood.

3.3 FATP1 Augments the Efficient Uptake of LCFAs via a Constitutive Interaction with ACSL1in Adipocytes

The FATP family comprises 6 highly homologous FA transport proteins in human/mouse/rat [41], also known as solute carrier protein family 27A (SLC27A) [42, 43]. FATP1, originally called FATP, the first of these proteins, was identified by Schaffer and Lodish in 1994 with an expression cloning strategy and a cDNA library from 3T3-L1 adipocytes. FATP1 is localized to the plasma membrane and augments the uptake of LCFAs when expressed in cultured cells [44]. FATP1 has a distinct membrane topology, and its N-terminus lies outside of the plasma membrane. Amino acids 1-190 contains at least 1 transmembrane domain, the fragment from 190–257 contains the AMP-binding motif, and the C-terminus faces the cytosolic space [45, 46]. FATP1-6 are expressed in a variety of tissues and, as membrane proteins, are associated with the import of FAs [41, 47, 48].

In addition to promoting cellular FA uptake, FATP has acyl-coA synthase activity that is central in downstream metabolic pathways, and overexpression of FATP1 increased acyl-CoA synthetase activity and fatty acid uptake in 3T3-L1 adipocytes [17, 49]. A constitutive interaction between FATP1 and A ligases/acyl-CoA synthetase 1 (ACSL1) contributes to the efficient cellular uptake of LCFAs in adipocytes [48, 50]. FATP1 and FATP4 localize to the endoplasmic reticulum to facilitate uptake and utilization of LCFAs by catalyzing the esterification of FAs with CoA [48, 49, 51].

3.4 Caveolin-1 Mediate the Translocation of LCFAs as the Major Structural Protein of Caveolae

Caveolae, specialized rafts, are 50- to 100-nm flask-shaped invaginations of the cell surface plasma membrane that are found in many cell types; transcytosis is one of the first functions of caveolae that transport macromolecules into the cell [52, 53]. Caveolae are associated with signal transduction and endocytosis of pathogens [54]. Pohl et al. [55] found a significant function for caveolae-mediated uptake and intracellular trafficking of LCFAs in HepG2 cells.

The biogenesis and function of caveolae depend on caveolins, of which there are 3 in mammals: Cav-1, -2, and -3 [53]. Caveolin-1 is the principal marker and major structural protein of caveolae, and its loss results in a complete lack of caveolae from the plasma membrane and influences FA uptake by regulating the availability of FAT/CD36 at the surface [56, 57]. Cav-1 can bind FAs directly [58], and the transport of FAs across the plasma membrane is modulated by caveolin-1 [59, 60]. Further, caveolin-1 accumulates LCFAs on the inner leaflet and presents them to cytoplasmic FABPs for further shuttling of LCFAs to various organelles [61].

3.5 “Flip-Flop” Mechanism Moves LCFAs in a Passive Diffusion Manner

Flip-flop. In addition to the 4 proteins above that help transport LCFAs across the plasma membrane, passive diffusion moves LCFAs from one half of the bilayer to the other—through the so-called “flip-flop” mechanism of diffusion (Table 2, Ref. [32, 33, 38, 39, 44, 53, 55, 62, 63]) [62, 63]. In early research on FA transport, passive diffusion was suggested and proven in a protein-free model of the phospholipid bilayer, and it was believed that this mechanism was efficient for shuttling FAs through simple membrane models [64]. However, this mode of LCFA transport across cell membranes remained controversial until now, and the focus has shifted to transport efficiency [65].

Table 2.Proteins in cellular uptake of LCFAs.
Name Function Molecular mass Mechanism Reference
FAT/CD36 Protein mediated facilitated diffusion 78–88 kDa CD36 is expressed by the mammary glands and respond to LCFAs to improve milk lipid synthesis. [32, 33]
FABPpm Protein mediated facilitated diffusion 40–43 kDa The overexpression of FABPpm increases the rate of LCFA transport and metabolism. [38, 39]
FATP Protein mediated facilitated diffusion 63 kDa FATP1 augments the uptake of LCFAs when expressed in cultured cells. [44]
Caveolin-1 Protein mediated facilitated diffusion 17–24 kDa Function for caveolae-mediated uptake and intracellular trafficcking of LCFAs. [53, 55]
Flippases, floppases and scramblases Passive diffusion Passive diffusion moves LCFAs from one half of the bilayer to the other—through the so-called “flip-flop” mechanism of diffusion. [62, 63]
FAT/CD36, FA translocase; FABPpm, plasma membrane FA-binding protein; FATP, FA transport protein; LCFA, Long-chain fatty acid; flippases, P4-ATPases; Floppases, ATP-binding cassette (ABC) transporters, ATP-Binding Cassette A1 (ABCA1); Scramblases, the TMEM16 family (anoctamins).

The free fatty acid transport model through the cell membrane is divided into 3 steps: adsorption of the FA to the membrane, translocation across the membrane (“flip-flop”), and subsequent desorption of the fatty acid into the cytosol [63, 66]. Several groups have suggested that the “flip-flop” step is not limiting—that desorption is the rate-limiting step in the phospholipid bilayer [67, 68]—and that FAs can diffuse freely by flip-flopping, even in the biological membrane [69, 70]. In contrast, other studies have suggested that the flip-flop of LCFAs is prohibitively slow and that transbilayer flip-flopping is rate-limiting in lipid bilayer membranes—that flip-flop across the lipid phase alone is unable to support the metabolic requirements of cells [66].

Although the debate regarding whether this process is fast or slow continues, many proteins that are associated with flip-flop have been identified [71, 72], falling into 3 broad categories: flippases, floppases, and scramblases; the first 2 groups are ATP-dependent, whereas scramblases facilitate the bidirectional movement of lipids in an ATP-independent manner [73]. Type 4 P-type ATPases (P4-ATPases) are flippases that mediate out-to-in lipid movement through the plasma membrane, from the exoplasmic to cytosolic side [74]. Floppases are transmembrane ATP-binding cassette (ABC) transporters that use the hydrolysis of ATP to facilitate the in-to-out movement of various substrates across the cell membrane [75]. ATP-Binding Cassette A1 (ABCA1) flops cholesterol from the inner to the outer leaflet of the plasma membrane [76]. The TMEM16 family of proteins, also known as anoctamins, contains lipid scramblases that are activated by increases in intracellular Ca2+ [77], and scramblases translocate phospholipids between the outer and inner leaflets of the cell membrane [71].

The lipid flip-flop model (passive diffusion) has been challenged by the discovery of several transporters that mediate LCFA translocation across the plasma membrane [27, 36, 44], generating alternative hypotheses, such as passive diffusion and protein-mediated FA transport through membranes (Fig. 2). It is likely that membrane proteins are important in the transmembrane transport of FAs across the plasma membrane.

Fig. 2.

Cellular uptake of LCFAs. Non-esterified LCFAs in blood exist as complexes with serum albumin. The cellular uptake of exogenous LCFAs requires the dissociation of FAs from albumin-FA complexes. Free LCFAs enter cells through passive diffusion and protein-mediated translocation across the plasma membrane. The free FA transport model alone, through a “flip-flop” mechanism across the lipid phase, is unable to support metabolic requirements, and the membrane proteins FAT/CD36, FABPpm, FATP, and caveolin-1 play key roles in LCFA translocation across the plasma membrane. The proteins associated with “flip-flop” fall into three broad categories: flippases, floppases, and scramblases, the first 2 of which are ATP-dependent, whereas scramblases facilitate the bidirectional movement of lipids in an ATP-independent manner. Membrane proteins play key roles in the transmembrane transport of LCFAs across the plasma membrane. ABC-transporters, ATP-binding cassette transporters; ACSL, A ligases/acyl-CoA synthetase; ER, endoplasmic reticulum; FAT/CD36, FA translocase; FABPpm, plasma membrane FA-binding protein; FATP, FA transport protein; LCFA, Long-chain fatty acid; P4-ATPases, Type 4 P-type ATPases; SLC27A, solute carrier protein family 27A.

4. Intracellular LCFA Metabolic Pathways

LCFAs are released from the inner leaflet of the plasma membrane into the cytoplasm after translocation across the plasma membrane by passive diffusion or protein-mediated diffusion [17]. The released FAs bind to cytoplasmic FA-binding proteins (FABPs), which are LCFA carriers with high cytosolic concentration [17, 78] to transport LCFAs to sites of metabolic conversion (e.g., oxidation, esterification) or subcellular targets [79, 80]. Cytoplasmic FABPs comprise a family of proteins that bind LCFAs with high affinity [81, 82], and at least 9 FABPs have been identified in human [82, 83]. FABPs have tissue-specific expression patterns and abound in tissues with active FA metabolism [84, 85].

The primary function of FABP family members is LCFA intracellular transport; they might also promote LCFA desorption from the cytoplasmic face of the plasma membrane [81]. Further, desorbed LCFAs are bound by FABP and transported to various organelles for metabolism, such as mitochondria for β-oxidation [79, 86] or assembly into peroxisomes for β-oxidation [87], and synthesis of triacylglycerol and other complex lipids on the outer leaflet of the endoplasmic reticulum [82, 88]. Non-esterified free FAs (NEFAs) must be activated to acyl-CoA by their acyl-CoA synthetase, as the first step of any metabolic process [89]. Long-chain fatty acid-coenzyme A ligases/acyl-CoA synthetases (ACSLs) have a significant function in activating LCFAs in mammals, and membrane FATP (SLC27A) proteins have acyl-coA synthase activity [49, 90, 91]. The process by which LCFAs are transported or activated into acyl-CoA by cytoplasmic proteins enables LCFAs or acyl-CoA to target specific metabolic pathways.

In this review, we discuss the main metabolic pathways of palmitic acid (PA, C16:0) and stearic acid (SA, C18:0) in cells as examples of LCFAs metabolism (Fig. 3). Cytoplasmic non-esterified free palmitic acid or stearic acid is activated to palmityl-CoA or stearoyl-CoA by ACSLs [92, 93] and then used in various metabolic pathways. Metabolically, intracellular PA and SA primarily undergo esterification, β-oxidation, desaturation, and elongation.

Fig. 3.

Intracellular LCFA metabolic pathways. Cytoplasmic non-esterified free palmitic acid (PA, C16:0) or stearic acid (SA, C18:0) must be activated to acyl-CoA by their acyl-CoA synthetase to enable them to target specific metabolic pathways—primarily esterification, β-oxidation, desaturation, and elongation to produce phospholipids, triacylglycerol, cholesteryl esters and special lipids, energy, VLCFAs, ULCFAs, MUFAs, and PUFAs. ACSL, A ligases/acyl-CoA synthetase; GPAT, glycerol-3-phosphate acyltransferase; AGPAT, 1-acylglycerol-3-phosphate O-acyltransferase; DGAT, diacylglycerol acyltransferase; FADS2, Fatty acid desaturase 2; SCD, stearoyl-CoA desaturase; TG synthesis, triacylglycerol synthesis; ELOVL, elongation of very long chain FAs; HACD, 3-hydroxyacyl-CoA dehydratase; KAR, 3-ketoacyl-CoA reductase; TER, trans-2, 3-enoyl-CoA reductase.

Esterification is the chief means by which FAs are used as cellular energy stores. After entering cells, non-esterified free LCFAs are activated by their acyl-CoA synthetase and targeted to phospholipids, triacylglycerol, or cholesteryl esters by FABP. NEFAs are preferentially stored as triacylglycerol via the sequential activities of glycerol-3-phosphate acyltransferase (GPAT), 1-acylglycerol-3-phosphate O-acyltransferase (AGPAT), PAPase/Lipin 1, and diacylglycerol acyltransferase (DGAT) [94, 95]. The triacylglycerol structure and interesterification of palmitic and stearic acids likely affect the physical characteristics of fat [96, 97] and their oxidative stability [98]. Recent reports revealed the crystal structure of the lipin/Pah phosphatidic acid phosphatase, suggesting its mechanisms [99], wherein the middle lipin domain adopts a novel membrane-binding dimeric protein fold, with which the functions of lipin oligomerization can be determine in the regulation of phospholipid and triacylglycerol synthesis [100]. In addition, LCFAs stored in phospholipids are presented in the plasma membrane. The arachidonic acid (AA), a polyunsaturated 20 carbon fatty acid, is released when the membrane phospholipids were hydrolyzed by phospholipase A2 (PLA2). AA can be subsequently metabolized by enzymes to generate prostaglandins (PGs) and other productions of the eicosanoid, which are involved in the regulation of many biological processes [101].

β-oxidation is the main pathway of LCFA catabolism, which occurs in the mitochondria and peroxisome [102]. Mitochondrial FA β-oxidation is a major catabolic process that degrades LCFAs [103]. Long-chain fatty acyl-CoA is needed for it to conjugate to carnitine, forming acylcarnitine, to enter the mitochondria. This action is catalyzed by carnitine acyl transferase. Acylcarnitine is transported by carnitine acylcarnitine translocase through the inner mitochondrial membrane [104, 105]. Long-chain fatty acyl-CoA that enters mitochondria completes β-oxidation in 4 steps, producing acetyl-CoA and acyl-CoA that shorten the fatty chain by 2 Cs. β-oxidation comprises oxidation, hydration, oxidation, and cleavage, which are catalyzed by acyl-CoA dehydrogenase, enoyl-acyl-CoA hydratase, β-hydroxy acyl-CoA dehydrogenase, and β-keto acyl-CoA-thiolase, respectively [104, 106]. Oxidation of FAs occurs in many areas in the cell. In addition, β-oxidation occurs in the mitochondria and peroxisome, whereas alpha-oxidation and omega-oxidation take place in the peroxisome and endoplasmic reticulum, respectively [107]. FA oxidation is a highly energetic process.

Desaturation or elongation converts LCFAs into other types of FAs. The desaturation of LCFAs is ubiquitous—e.g., the saturated stearic acid (C18:0) is converted to monounsaturated oleic acid (C18:1n-9) by stearoyl CoA desaturase (SCD) [108]. SCD has 2 isoforms (SCD1 and SCD5) in humans and catalyzes Δ9-desaturation to biosynthesize MUFAs, particularly oleic acid (18:1n-9) and palmitoleic acid (16:1n-7) [109]. Another type of desaturase is the FA desaturase (FADS) family, which consists of FADS1, FADS2, and FADS3; palmitic acid 16:0 is preferentially converted to sapienic acid 16:1n-10 rather than 16:1n-7 by FADS2 [109, 110]. SCD-dependent FA desaturation is considered to be the only source of de novo-generated monounsaturated FAs, and FAD2 can desaturate palmitate to the atypical FA sapienate in cancer cells [111]. The dual pathway of SCD- and FADS2-dependent desaturation can provide plasticity in saturated fatty acid metabolism.

LCFAs can be used to synthesize membrane or other phospholipids, for which LCFAs are often elongated [111, 112]. In metabolic experiments with [1-14C]-labeled myristic acid (C14:0) and palmitic acid (16:0), myristic acid was strongly elongated to radiolabeled palmitic acid, and palmitic acid was lengthened to stearic acid [113]. Palmitic acid (16:0), which is synthesized de novo and taken up from the diet, can be elongated into stearic acid (C18:0) and further to VLCFAs. Formation of VLCFAs is performed mainly in the endoplasmic reticulum by membrane-bound enzymes [114], and elongation occurs by cycling through a 4-step process (condensation, reduction, dehydration, and reduction), corresponding to 4 enzymes: elongases (elongation of very long chain FAs, ELOVL), 3-ketoacyl-CoA reductase (KAR), 3-hydroxyacyl-CoA dehydratase (HACD), and trans-2, 3-enoyl-CoA reductase (TER) (Table 3, Ref. [94, 95, 102, 108, 109, 111, 112] ).

Table 3.The main metabolic pathways of palmitic acid and stearic acid.
Metabolic pathways Function Products Reference
Esterification The chief means by which FAs are used to form triacylglycerol and complex lipids. phospholipid, triacylglycerol, or cholesteryl ester [94, 95]
β-oxidation The main pathway of LCFA catabolism, which occurs in the mitochondria and peroxisome. acetyl-CoA [102]
Desaturation Desaturation or elongation converts LCFAs into other types of FAs. C16:1, C18:1 [108, 109]
Elongation LCFAs can be used to synthesize membrane or other phospholipids, for which LCFAs are often elongated. stearoyl-CoA [111, 112]
FAs, Fatty acids; FADS2, Fatty acid desaturase 2; LCFAs, Long-chain fatty acids; SCD, stearoyl-CoA desaturase.

VLCFA elongation has been reviewed extensively [114]. The condensation that is catalyzed by elongase (ELOVL) is the rate-limiting step in the sequential VLCFA elongation cycle, and a recent study reported the first crystal structure of a membrane-bound FA elongase-condensing enzyme, revealing a new reaction mechanism: FA elongation by ELOVL depends on a histidine nucleophile [115, 116]. These novel crystal structures of lipid metabolism enzymes provide important insights into the reaction mechanism of LCFAs.

5. LCFA Sensing by mTORC1

Mammalian cells must adapt their metabolism to maintain their energy homoeostasis and respond to nutrient availability during their proliferation. Thus, the ability to properly sense both ingested and circulating nutrients is crucial for the maintenance of metabolic homeostasis [117]. Cellular nutrient sensing mechanisms engage anabolism and storage when food abundance, and scarcity triggers homeostatic mechanisms, such as uptake from outside the cell or the mobilization of internal stores [118].

Mammalian target of rapamycin (mTOR), which is now referred to as mechanistic target of rapamycin, has been implicated as a sensor of nutrient sufficiency in cells and is activated by essential amino acids, glucose, and phosphatidic acid (PA) [119, 120, 121]. mTOR is a kind of Ser/Thr kinase in mammalian cells, and it combines with other proteins to form two mTOR complexes, mTORC1 and mTORC2. mTORC1 integrates input signals from nutrients, growth factors, energy, oxygen and environmental stress to control cellular growth and metabolism homeostasis [122]. Nutrients, including amino acids, glucose and nucleotide, drive the recruitment of mTORC1 to the lysosomal surface via the Rag GTPases and the sensors of several types of amino acid, glucose and purines in cells mediated nutrient signals to excite mTORC1 signaling are reported in recent years [122, 123]. mTORC1 is also involved in lipid metabolism and several types of fatty acid signals can excite mTORC1 [124, 125, 126, 127]. However, the sensing mechanisms of fatty acid in cells and how mTORC1 are regulated by fatty acid signals remain unclear. Therefore, understanding the cellular mechanism behind the fatty acid sensing and homeostasis is important for cell growth and metabolism.

The fluctuation of internal fatty acid levels and intracellular and extracellular LCFA sensing mechanisms exist in mammals. 2 types of pathway may exist in mammalian cells, they are the direct binding of the LCFA molecule to the sensor protein, or indirect mechanism relying on the detection of metabolites of LCFA [118]. Extracellular and intracellular LCFAs may sense by sensors which located on cytoplasmic membrane and in cytoplasm. Increasing evidence in human or rodent models indicates that G protein-coupled receptors (GPRs), as free fatty acid receptors (FFARs), can sense the level of extracellular LCFA and affect the biological characteristics of cells. The free fatty acid receptors include FFAR1 (GPR40), FFAR2 (GPR43), FFAR3 (GPR41) and FFAR4 (GPR120) which are seven transmembrane-spanning proteins show different expression patterns on different cells, and GPR40 (FFAR1) and GPR120 (FFAR4) are involved in sensing medium- and long-chain fatty acids while GPR43 (FFAR2) and GPR41 (FFAR3) are activated by SCFA [17, 128]. In addition to GPR40 and GPR120, other GPR-independent mechanisms have been suggested in mediating LCFA sensing, e.g., CD36 [117, 118, 129]. CD36, as mentioned earlier, is a multifunctional membrane protein, and is considered as the most important translocator of LCFAs. Importantly, CD36 has been reported to be a LCFA receptor and involved in LCFA sensing [130, 131, 132, 133].

GPRs couple to diverse intracellular downstream G proteins and then activate downstream signaling pathways. In general, GPR40 or GPR120 are activated by the extracellular LCFAs, and then transduce signals downstream cAMP and the phospholipase C (PLC) signaling cascade. PLC cleaves the membrane lipid phosphatidylinositol 4,5-bisphosphate (PIP2) into the second messengers diacylgycerol (DAG) and inositol 1,4,5-trisphosphate (IP3), leading to calcium release, protein kinase C (PKC) activation, and the phosphatidylinositol 3-kinase (PI3K)/Akt (protein kinase B) signaling pathway [134, 135]. BSA-conjugated palmitic acid increases Akt/mTORC1 pathway via GPR40, and the mechanism by which palmitic acid regulate mTORC1 activity is probably its translocation onto lysosomal membranes [136, 137]. Moreover, oleic acid activates Akt/mTORC1 and ERK/mTORC1 pathways via GPR40 or GPR120 [138], and inhibits AMPK signaling which is a negative regulator of mTORC1 [139, 140, 141, 142]. There are few reports on stearic acid and mTORC1 activity, it can upregulate mTOR expression [143].

LCFA, e.g., palmitic acid, can be used to synthesize phosphatidic acid (PA) or be degraded to produce acetyl-CoA after LCFAs are transported inside cells. Phosphatidic acid interacts with the FK506-binding protein–12-rapamycin-binding (FRB) domain of mTOR to activate mTORC1 [120, 144, 145], and lower total cytosolic acetyl-CoA levels led to decreased raptor acetylation and reduced lysosomal localization of mTOR, resulting in impaired activation of mTORC1 [126], indicating changes in the level of LCFA metabolites can be sensed by mTORC1. However, whether there is a mechanism for mTOR to distinguish phosphatidic acid species with LCFAs remains to be determined (Table 4, Ref. [18, 134, 138, 141, 144]).

Table 4.Fatty acid receptors and involved in LCFA sensing.
Protein name FA type Signaling pathway mTORC1 signaling Reference
GPR40 MCFA and LCFA (a) cAMP/PKA/ERK/mTORC1 GPR40 mediates extracellular LCFA signals to excite mTORC1 signaling [18, 134]
(b) PLC/DAG/PKC/ERK/mTORC1
GPR120 MCFA and LCFA (a) PLC/IP3/AMPK/mTORC1 GPR120 mediates extracellular LCFA signals to excite mTORC1 signaling [138, 141]
(b) PI3K/Akt/mTORC1
CD36 LCFA (a) PA/ mTORC1 Intracellular LCFAs translocated by CD36 are further metabolized to produce PA or acetyl-CoA, which regulates mTORC1 activity [144]
AMPK, AMP-activated protein kinase; Akt, Protein kinase B; cAMP, Cyclic adenosine monophosphate; DAG, Diacylgycerol; ERK, Extracellular signal-regulated kinase; GPR40, G protein-coupled receptor 40; GPR120, G protein-coupled receptor 120; IP3,Inositol 1,4,5-trisphosphate; LCFA, Long-chain fatty acid; MCFA, Medium-chain fatty acid; mTORC1, Mammalian target of rapamycin complex 1; PKA, Protein kinase A system; PLC, Phospholipase C; PKC, Protein kinase C; PI3K, Phosphatidylinositol 3-kinase; PA, Phosphatidic acid.

The downstream signaling pathways mediated by GPR40 or GPR120, the second messenger cAMP can activate downstream effector PKA and regulate ERK sequentially [138, 146, 147], and another downstream pathway PLC/DAG/IP3 in which DAG can excite PKC and ERK/mTORC1 [147, 148], and IP3 leads to calcium release activating AMPK which is an upstream negative regulator of mTORC1 [139, 140, 142, 149], and the third downstream pathway, PI3K/Akt pathway, can activate mTORC1 signaling [136, 138, 140, 147]. Different G proteins mediate LCFAs to regulate mTORC1 signaling. Furthermore, although whether CD36 can act as a LCFAs receptor to activate mTORC1 remains determined, and it was definite that intracellular LCFAs translocated by CD36 are further metabolized to produce phosphatidic acid (PA) or acetyl-CoA, which regulates mTORC1 activity (Fig. 4). Whether mTORC1 senses intracellular LCFA levels or existing other sensors of LCFAs remains to be determined.

Fig. 4.

LCFAs sensing by mTORC1. The downstream signaling pathways of GPR40 and GPR120, including cAMP, PLC/DAG/IP3, and PI3K/Akt. cAMP activates downstream effector PKA and regulate ERK sequentially; PLC/DAG/IP3 excites PKC and ERK/mTORC1; IP3 leads to calcium release activating AMPK/mTORC1; PI3K/Akt activates mTORC1 signaling. Phosphatidic acid (PA) or acetyl-CoA (AC) regulates mTORC1 activity. Whether mTORC1 can directly sense intracellular LCFA levels remains to be determined. AC, acetyl-CoA; AMPK, AMP-activated protein kinase; Akt, Protein kinase B; cAMP, Cyclic adenosine monophosphate; DAG, Diacylgycerol; ERK, Extracellular signal-regulated kinase; GPR40, G protein-coupled receptor 40; GPR120, G protein-coupled receptor 120; IP3, Inositol 1,4,5-trisphosphate; LCFA, Long-chain fatty acid; mTORC1, Mammalian target of rapamycin complex 1; PKA, Protein kinase A system; PLC, Phospholipase C; PKC, Protein kinase C; PI3K, Phosphatidylinositol 3-kinase; PA, Phosphatidic acid.

6. Conclusions

In this review, we have provided the most up-to-date information available on the absorption of LCFAs, their translocation across the plasma membrane, and their metabolic pathways as well as their sensing mechanisms in cells. We have not covered all possible metabolic pathways and regulatory mechanisms of LCFA, such as the regulation of the transcriptome by fatty acids, or whether mTOR to distinguish phosphatidic acid species with different types LCFAs in cells. However, based on the literature reviewed, we can make some concluding comments and propose some forward-looking predictions for each main topic covered in this review.

FAs are important biocompounds that participate in complex metabolic pathways to ensure human health and development. FAs are obtained by mammalian cells through external uptake (primarily), de novo synthesis, hydrolysis of triacylglycerols. LCFAs and VLCFAs are absorbed mainly from outside of cells in mammals. LCFAs enter cells through passive diffusion and protein-mediated FA translocation across the plasma membrane, in the latter of which FAT/CD36, FABPpm, FATP, and caveolin-1 are believed to be critical. The LCFAs that are absorbed by cells bind to FABPs, are transported to metabolic organelles, and converted into acyl-CoA to target specific metabolic pathways. LCFA-CoA is esterified to phospholipids, triacylglycerols, and cholesteryl esters or other specific lipids.

NEFAs are preferentially stored as triacylglycerol, and the triacylglycerol structure and interesterification of LCFAs likely affect the physical characteristics of fat and oxidative stability. The mechanisms by which triacylglycerol synthase function will be illustrated by the crystal structure of these synthases. β-oxidation, which occurs in mitochondria, has been studied extensively, but other types of FA oxidation in the peroxisome, endoplasmic reticulum, and other subcellular structures are poorly understood. SCD-dependent FA desaturation is well characterized, unlike FADS-dependent desaturation, and the metabolic significance of this dual activity needs to be further examined.

FA elongation occurs by cycling through a 4-step process, comprising condensation, reduction, dehydration, and reduction. The first crystal structure of a membrane-bound FA elongase-condensing enzyme has revealed a new reaction mechanism. Understanding the mechanism of the cellular uptake and metabolism of FAs is important for human nutrition and metabolism, and the novel crystal structures of lipid metabolism enzymes provide insights into the reaction mechanism of LCFAs.

GPR40 or GPR120 mediates extracellular LCFA signals to excite mTORC1 signaling, and intracellular LCFA’s sensor remains to be determined. However, mTORC1 activation can be regulated by phosphatidic acid (PA) and acetyl-CoA which are the metabolites of LCFAs. CD36 is a potential receptor of LCFAs to mediate extracellular LCFA signals.

The metabolism of LCFAs is closely related to the occurrence of human obesity, nonalcoholic fatty liver disease (NAFLD), cardiovascular disease, and hyperlipidemia. The metabolism of LCFAs is diverse and is related to genes, living environment and dietary nutrition. Thus, the metabolism of different LCFAs, and the regulatory mechanism in human health and diseases should be considered comprehensively.

Abbreviations

ABC-transporters, ATP-binding cassette transporters; ACSL, A ligases/acyl-CoA synthetases; AMPK, AMP-activated protein kinase; Akt, Protein kinase B; cAMP, Cyclic adenosine monophosphate; DG/DAG, Diacylgycerol; PKC, Protein kinase C; ERK, Extracellular signal-regulated kinase; ER, endoplasmic reticulum; FAs, Fatty Acids; FAT/CD36, FA translocase; FABPpm, plasma membrane FA-binding protein; FATP, FA transport protein; FATP1 originally called FATP; FADS2, Fatty acid desaturase 2; GPR40, G protein-coupled receptor 40; GPR120, G protein-coupled receptor 120; IP3, Inositol 1,4,5-trisphosphate; LCFAs, Long-chain fatty acids; mTORC1, Mammalian target of rapamycin complex 1; MCFAs, Medium-chain fatty acids; PA, Phosphatidic acid; PKA, Protein kinase A system; PLC, Phospholipase C; PI3K, Phosphatidylinositol 3-kinase; SCFAs, Short-chain fatty acids; SLC27A, Solute carrier protein family 27A; SCD, Stearoyl-CoA desaturase; TG/TAG, Triacylglycerol; VLCFAs, Very long-chain fatty acids.

Author Contributions

PY—writing - reviewing and editing, supervision. QH—writing - original draft reviewing and editing. YC—visualization, investigation. ZW—conceptualization, writing - reviewing and editing. HH—writing - reviewing and editing. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This work was funded by the Inner Mongolia Agriculture and Animal Husbandry Innovation Fund (2019CXJJM11 and 2022CXJJM09), by Inner Mongolia Autonomous Region Science and Technology Program Project (2019GG354).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Ulug E, Nergiz-Unal R. Dietary fatty acids and CD36-mediated cholesterol homeostasis: potential mechanisms. Nutrition Research Reviews. 2021; 34: 64–77.
[2]
Panaroni C, Fulzele K, Mori T, Siu KT, Onyewadume C, Maebius A, et al. Multiple myeloma cells induce lipolysis in adipocytes and uptake fatty acids through fatty acid transporter proteins. Blood. 2022; 139: 876–888.
[3]
Cao D, Song X, Che L, Li X, Pilo MG, Vidili G, et al. Both de novo synthetized and exogenous fatty acids support the growth of hepatocellular carcinoma cells. Liver International. 2017; 37: 80–89.
[4]
Stergiadis S, Cabeza-Luna I, Mora-Ortiz M, Stewart RD, Dewhurst RJ, Humphries DJ, et al. Unravelling the Role of Rumen Microbial Communities, Genes, and Activities on Milk Fatty Acid Profile Using a Combination of Omics Approaches. Frontiers in microbiology. 2020; 11: 590441.
[5]
Zhang ZD, Wang C, Du HS, Liu Q, Guo G, Huo WJ, et al. Effects of sodium selenite and coated sodium selenite on lactation performance, total tract nutrient digestion and rumen fermentation in Holstein dairy cows. Animal. 2020; 14: 2091–2099.
[6]
Costa LA, de Araujo MJ, Edvan RL, Bezerra LR, de Sousa AR, Viana FJC, et al. Chemical composition, fermentative characteristics, and in situ ruminal degradability of elephant grass silage containing Parkia platycephala pod meal and urea. Tropical Animal Health and Production. 2020; 52: 3481–3492.
[7]
Alhusseiny SM, El-Beshbishi SN. Omega polyunsaturated fatty acids and parasitic infections: An overview. Acta Tropica. 2020; 207: 105466.
[8]
Siziba LP, Lorenz L, Stahl B, Mank M, Marosvolgyi T, Decsi T, et al. Human Milk Fatty Acid Composition of Allergic and Non-Allergic Mothers: The Ulm SPATZ Health Study. Nutrients. 2020; 12: 1740.
[9]
Lee YY, Tang TK, Chan ES, Phuah ET, Lai OM, Tan CP, et al. Medium chain triacylglycerol and medium-and long chain triacylglycerol: metabolism, production, health impacts and its applications–a review. Critical Reviews in Food Science and Nutrition. 2022; 62: 4169–4185.
[10]
Sugiyama T, Hobro AJ, Pavillon N, Umakoshi T, Verma P, Smith N. Label-free Raman mapping of saturated and unsaturated fatty acid uptake, storage, and return toward baseline levels in macrophages. Analyst. 2021; 146: 1268–1280.
[11]
Ajie HO, Connor MJ, Lee WN, Bassilian S, Bergner EA, Byerley LO. In vivo study of the biosynthesis of long-chain fatty acids using deuterated water. American Journal of Physiology-Endocrinology and Metabolism. 1995; 269: E247–E252.
[12]
Mika A, Kobiela J, Pakiet A, Czumaj A, Sokolowska E, Makarewicz W, et al. Preferential uptake of polyunsaturated fatty acids by colorectal cancer cells. Scientific Reports. 2020; 10: 1–8.
[13]
Tvrzicka E, Kremmyda LS, Stankova B, Zak A. Fatty acids as biocompounds: their role in human metabolism, health and disease–a review. Part 1: classification, dietary sources and biological functions. Biomedical Papers of the Medical Faculty of Palacky University in Olomouc. 2011; 155: 117–130.
[14]
Guo Z, Wang Y, Feng X, Bao C, He Q, Bao L, et al. Rapamycin Inhibits Expression of Elongation of Very-long-chain Fatty Acids 1 and Synthesis of Docosahexaenoic Acid in Bovine Mammary Epithelial Cells. Asian-Australasian Journal of Animal Sciences. 2016; 29: 1646–1652.
[15]
Sassa T, Kihara A. Metabolism of very long-chain Fatty acids: genes and pathophysiology. Biomolecules & Therapeutics. 2014; 22: 83–92.
[16]
Benoit B, Bruno J, Kayal F, Estienne M, Debard C, Ducroc R, et al. Saturated and Unsaturated Fatty Acids Differently Modulate Colonic Goblet Cells In Vitro and in Rat Pups. The Journal of Nutrition. 2015; 145: 1754–1762.
[17]
Bionaz M, Vargas-Bello-Perez E, Busato S. Advances in fatty acids nutrition in dairy cows: from gut to cells and effects on performance. Journal of Animal Science and Biotechnology. 2020; 11: 110.
[18]
Kimura I, Ichimura A, Ohue-Kitano R, Igarashi M. Free Fatty Acid Receptors in Health and Disease. Physiological Reviews. 2020; 100: 171–210.
[19]
Koh A, De Vadder F, Kovatcheva-Datchary P, Backhed F. From Dietary Fiber to Host Physiology: Short-Chain Fatty Acids as Key Bacterial Metabolites. Cell. 2016; 165: 1332–1345.
[20]
Schönfeld P, Wojtczak L. Short- and medium-chain fatty acids in energy metabolism: the cellular perspective. Journal of Lipid Research. 2016; 57: 943–954.
[21]
Janssen CI, Kiliaan AJ. Long-chain polyunsaturated fatty acids (LCPUFA) from genesis to senescence: the influence of LCPUFA on neural development, aging, and neurodegeneration. Progress in Lipid Research. 2014; 53: 1–17.
[22]
Qi L, Yan S, Sheng R, Zhao Y, Guo X. Effects of Saturated Long-chain Fatty Acid on mRNA Expression of Genes Associated with Milk Fat and Protein Biosynthesis in Bovine Mammary Epithelial Cells. Asian-Australasian Journal of Animal Sciences. 2014; 27: 414–421.
[23]
Schweiger M, Eichmann TO, Taschler U, Zimmermann R, Zechner R, Lass A. Measurement of Lipolysis. Methods in Enzymology. 2014; 284: 171–193.
[24]
Luo J, Yang H, Song B. Mechanisms and regulation of cholesterol homeostasis. Nature Reviews Molecular Cell Biology. 2020; 21: 225–245.
[25]
McArthur MJ, Atshaves BP, Frolov A, Foxworth WD, Kier AB, Schroeder F. Cellular uptake and intracellular trafficking of long chain fatty acids. Journal of Lipid Research. 1999; 40: 1371–1383.
[26]
Cifarelli V, Abumrad NA. Intestinal CD36 and Other Key Proteins of Lipid Utilization: Role in Absorption and Gut Homeostasis. Comprehensive Physiology. 2018; 8: 493–507.
[27]
Abumrad NA, el-Maghrabi MR, Amri EZ, Lopez E, Grimaldi PA. Cloning of a rat adipocyte membrane protein implicated in binding or transport of long-chain fatty acids that is induced during preadipocyte differentiation. Homology with human CD36. Journal of Biological Chemistry. 1993; 268: 17665–17668.
[28]
Armesilla AL, Vega MA. Structural organization of the gene for human CD36 glycoprotein. The Journal of Biological Chemistry. 1994; 269: 18985–18991.
[29]
Luiken JJFP, Chanda D, Nabben M, Neumann D, Glatz JFC. Post-translational modifications of CD36 (SR-B2): Implications for regulation of myocellular fatty acid uptake. Biochimica Et Biophysica Acta (BBA) - Molecular Basis of Disease. 2016; 1862: 2253–2258.
[30]
Yang X, Okamura DM, Lu X, Chen Y, Moorhead J, Varghese Z, et al. CD36 in chronic kidney disease: novel insights and therapeutic opportunities. Nature Reviews Nephrology. 2017; 13: 769–781.
[31]
Ding Z, Liu S, Wang X, Theus S, Deng X, Fan Y, et al. PCSK9 regulates expression of scavenger receptors and ox-LDL uptake in macrophages. Cardiovascular Research. 2018; 114: 1145–1153.
[32]
Bernard L, Montazer Torbati MB, Graulet B, Leroux C, Chilliard Y. Long-chain fatty acids differentially alter lipogenesis in bovine and caprine mammary slices. Journal of Dairy Research. 2013; 80: 89–95.
[33]
Fougere H, Bernard L. Effect of diets supplemented with starch and corn oil, marine algae, or hydrogenated palm oil on mammary lipogenic gene expression in cows and goats: A comparative study. Journal of Dairy Science. 2019; 102: 768–779.
[34]
Hua W, Huang HZ, Tan LT, Wan JM, Gui HB, Zhao L, et al. CD36 Mediated Fatty Acid-Induced Podocyte Apoptosis via Oxidative Stress. PLoS ONE. 2015; 10: e0127507.
[35]
Holloway GP, Nickerson JG, Lally JSV, Petrick HL, Dennis KMJH, Jain SS, et al. Co-overexpression of CD36 and FABPpm increases fatty acid transport additively, not synergistically, within muscle. American journal of physiology. Cell Physiology. 2022; 322: C546–C553.
[36]
Schwieterman W, Sorrentino D, Potter BJ, Rand J, Kiang CL, Stump D, et al. Uptake of oleate by isolated rat adipocytes is mediated by a 40-kDa plasma membrane fatty acid binding protein closely related to that in liver and gut. Proceedings of the National Academy of Sciences of the United States of America. 1988; 85: 359–363.
[37]
Sorrentino D, Stump D, Potter BJ, Robinson RB, White R, Kiang CL, et al. Oleate uptake by cardiac myocytes is carrier mediated and involves a 40-kD plasma membrane fatty acid binding protein similar to that in liver, adipose tissue, and gut. The Journal of clinical investigation. 1988; 82: 928–935.
[38]
Isola LM, Zhou SL, Kiang CL, Stump DD, Bradbury MW, Berk PD. 3T3 fibroblasts transfected with a cDNA for mitochondrial aspartate aminotransferase express plasma membrane fatty acid-binding protein and saturable fatty acid uptake. Proceedings of the National Academy of Sciences. 1995; 92: 9866–9870.
[39]
Clarke DC, Miskovic D, Han XX, Calles-Escandon J, Glatz JF, Luiken JJ, et al. Overexpression of membrane-associated fatty acid binding protein (FABPpm) in vivo increases fatty acid sarcolemmal transport and metabolism. Physiological Genomics. 2004; 17: 31–37.
[40]
Wang TY, Liu M, Portincasa P, Wang DQ. New insights into the molecular mechanism of intestinal fatty acid absorption. European Journal of Clinical Investigation. 2013; 43: 1203–1223.
[41]
Hirsch D, Stahl A, Lodish HF. A family of fatty acid transporters conserved from mycobacterium to man. Proceedings of the National Academy of Sciences of the United States of America. 1998; 95: 8625–8629.
[42]
Zhang M, Di Martino JS, Bowman RL, Campbell NR, Baksh SC, Simon-Vermot T, et al. Adipocyte-Derived Lipids Mediate Melanoma Progression via FATP Proteins. Cancer Discovery. 2018; 8: 1006–1025.
[43]
Anderson CM, Stahl A. SLC27 fatty acid transport proteins. Molecular Aspects of Medicine. 2013; 34: 516–528.
[44]
Schaffer JE, Lodish HF. Expression cloning and characterization of a novel adipocyte long chain fatty acid transport protein. Cell. 1994; 79: 427–436.
[45]
Lewis SE, Listenberger LL, Ory DS, Schaffer JE. Membrane topology of the murine fatty acid transport protein 1. The Journal of Biological Chemistry. 2001; 276: 37042–37050.
[46]
Schaffer JE. Fatty acid transport: the roads taken. American journal of physiology. Endocrinology and Metabolism. 2002; 282: E239–246.
[47]
Pohl J, Ring A, Hermann T, Stremmel W. Role of FATP in parenchymal cell fatty acid uptake. Biochim Biophys Acta. 2004; 1686: 1–6.
[48]
Dourlen P, Sujkowski A, Wessells R, Mollereau B. Fatty acid transport proteins in disease: New insights from invertebrate models. Progress in Lipid Research. 2015; 60: 30–40.
[49]
Zhan T, Poppelreuther M, Ehehalt R, Füllekrug J. Overexpressed FATP1, ACSVL4/FATP4 and ACSL1 increase the cellular fatty acid uptake of 3T3-L1 adipocytes but are localized on intracellular membranes. PLoS ONE. 2012; 7: e45087.
[50]
Richards MR, Harp JD, Ory DS, Schaffer JE. Fatty acid transport protein 1 and long-chain acyl coenzyme A synthetase 1 interact in adipocytes. Journal of Lipid Research. 2006; 47: 665–672.
[51]
Huang J, Zhu R, Shi D. The role of FATP1 in lipid accumulation: a review. Molecular and Cellular Biochemistry. 2021; 476: 1897–1903.
[52]
Frank PG, Woodman SE, Park DS, Lisanti MP. Caveolin, caveolae, and endothelial cell function. Arteriosclerosis, Thrombosis, and Vascular Biology. 2003; 23: 1161–1168.
[53]
Parton RG. Caveolae: Structure, Function, and Relationship to Disease. Annual Review of Cell and Developmental Biology. 2018; 34: 111–136.
[54]
Stenqvist J, Carlsson T, Winder M, Aronsson P. Effects of caveolae depletion and urothelial denudation on purinergic and cholinergic signaling in healthy and cyclophosphamide-induced cystitis in the rat bladder. Autonomic Neuroscience. 2018; 213: 60–70.
[55]
Pohl J, Ring A, Stremmel W. Uptake of long-chain fatty acids in HepG2 cells involves caveolae: analysis of a novel pathway. Journal of Lipid Research. 2002; 43: 1390–1399.
[56]
Drab M, Verkade P, Elger M, Kasper M, Lohn M, Lauterbach B, et al. Loss of caveolae, vascular dysfunction, and pulmonary defects in caveolin-1 gene-disrupted mice. Science. 2001; 293: 2449–2452.
[57]
Ring A, Le Lay S, Pohl J, Verkade P, Stremmel W. Caveolin-1 is required for fatty acid translocase (FAT/CD36) localization and function at the plasma membrane of mouse embryonic fibroblasts. Biochimica et Biophysica Acta. 2006; 1761: 416–423.
[58]
Trigatti BL, Anderson RGW, Gerber GE. Identification of Caveolin-1 as a Fatty Acid Binding Protein. Biochemical and Biophysical Research Communications. 1999; 255: 34–39.
[59]
Meshulam T, Simard JR, Wharton J, Hamilton JA, Pilch PF. Role of caveolin-1 and cholesterol in transmembrane fatty acid movement. Biochemistry. 2006; 45: 2882–2893.
[60]
Otis JP, Shen MC, Quinlivan V, Anderson JL, Farber SA. Intestinal epithelial cell caveolin 1 regulates fatty acid and lipoprotein cholesterol plasma levels. Disease Models & Mechanisms. 2017; 10: 283–295.
[61]
Pohl J, Ring A, Ehehalt R, Herrmann T, Stremmel W. New concepts of cellular fatty acid uptake: role of fatty acid transport proteins and of caveolae. The Proceedings of the Nutrition Society. 2004; 63: 259–262.
[62]
Kleinfeld AM. Lipid phase fatty acid flip-flop, is it fast enough for cellular transport. The Journal of Membrane Biology. 2000; 175: 79–86.
[63]
Kamp F, Hamilton JA. How fatty acids of different chain length enter and leave cells by free diffusion. Prostaglandins Leukot Essent Fatty Acids. 2006; 75: 149–159.
[64]
Zhang F, Kamp F, Hamilton JA. Dissociation of long and very long chain fatty acids from phospholipid bilayers. Biochemistry. 1996; 35: 16055–16060.
[65]
Carley AN, Kleinfeld AM. Flip-flop is the rate-limiting step for transport of free fatty acids across lipid vesicle membranes. Biochemistry. 2009; 48: 10437–10445.
[66]
Cheng V, Kimball DR, Conboy DJC. Determination of the Rate-Limiting Step in Fatty Acid Transport. The Journal of Physical Chemistry. 2019; 123: 7157–7168.
[67]
Simard JR, Pillai BK, Hamilton JA. Fatty acid flip-flop in a model membrane is faster than desorption into the aqueous phase. Biochemistry. 2008; 47: 9081–9089.
[68]
Wei C, Pohorille A. Flip-flop of oleic acid in a phospholipid membrane: rate and mechanism. The Journal of Physical Chemistry B. 2014; 118: 12919–12926.
[69]
Hamilton JA, Civelek VN, Kamp F, Tornheim K, Corkey BE. Changes in internal pH caused by movement of fatty acids into and out of clonal pancreatic beta-cells (HIT). Journal of Biological Chemistry. 1994; 269: 20852–20856.
[70]
Kamp F, Guo W, Souto R, Pilch PF, Corkey BE, Hamilton JA. Rapid flip-flop of oleic acid across the plasma membrane of adipocytes. The Journal of Biological Chemistry. 2003; 278: 7988–7995.
[71]
Ikeda M, Kihara A, Igarashi Y. Lipid asymmetry of the eukaryotic plasma membrane: functions and related enzymes. Biological & Pharmaceutical Bulletin. 2006; 29: 1542–1546.
[72]
Contreras FX, Sanchez-Magraner L, Alonso A, Goni FM. Transbilayer (flip-flop) lipid motion and lipid scrambling in membranes. FEBS Letters. 2010; 584: 1779–1786.
[73]
Kobayashi T, Menon AK. Transbilayer lipid asymmetry. Current Biology. 2018; 28: R386–R391.
[74]
Timcenko M, Lyons JA, Januliene D, Ulstrup JJ, Dieudonné T, Montigny C, et al. Structure and autoregulation of a P4-ATPase lipid flippase. Nature. 2019; 571: 366–370.
[75]
Tarling EJ, de Aguiar Vallim TQ, Edwards PA. Role of ABC transporters in lipid transport and human disease. Trends in Endocrinology & Metabolism. 2013; 24: 342–350.
[76]
Okamoto Y, Tomioka M, Ogasawara F, Nagaiwa K, Kimura Y, Kioka N, et al. C-terminal of ABCA1 separately regulates cholesterol floppase activity and cholesterol efflux activity. Bioscience, Biotechnology, and Biochemistry. 2020; 84: 764–773.
[77]
Brunner JD, Lim NK, Schenck S, Duerst A, Dutzler R. X-ray structure of a calcium-activated TMEM16 lipid scramblase. Nature. 2014; 516: 207–212.
[78]
Thumser AEA, Storch J. Liver and intestinal fatty acid-binding proteins obtain fatty acids from phospholipid membranes by different mechanisms. Journal of Lipid Research. 2000; 41: 647–656.
[79]
Storch J, Thumser AEA. The fatty acid transport function of fatty acid-binding proteins. Biochimica Et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 2000; 1486: 28–44.
[80]
Schroeder F, Petrescu AD, Huang H, Atshaves BP, McIntosh AL, Martin GG, et al. Role of fatty acid binding proteins and long chain fatty acids in modulating nuclear receptors and gene transcription. Lipids. 2008; 43: 1–17.
[81]
Vork MM, Glatz JF, Van Der Vusse GJ. On the mechanism of long chain fatty acid transport in cardiomyocytes as facilitated by cytoplasmic fatty acid-binding protein. Journal of Theoretical Biology. 1993; 160: 207-222.
[82]
Smathers RL, Petersen DR. The human fatty acid-binding protein family: Evolutionary divergences and functions. Human Genomics. 2011; 5: 170.
[83]
Chmurzyńska A. The multigene family of fatty acid-binding proteins (FABPs): Function, structure and polymorphism. Journal of Applied Genetics. 2006; 47: 39–48.
[84]
Richieri GV, Ogata RT, Zimmerman AW, Veerkamp JH, Kleinfeld AM. Fatty acid binding proteins from different tissues show distinct patterns of fatty acid interactions. Biochemistry. 2000; 39: 7197–7204.
[85]
Wang L, Li L, Jiang J, Wang Y, Zhong T, Chen Y, et al. Molecular characterization and different expression patterns of the FABP gene family during goat skeletal muscle development. Molecular Biology Reports. 2015; 42: 201–207.
[86]
Sayed-Ahmed MM, Aldelemy ML, Al-Shabanah OA, Hafez MM, Al-Hosaini KA, Al-Harbi NO, et al. Inhibition of gene expression of carnitine palmitoyltransferase I and heart fatty acid binding protein in cyclophosphamide and ifosfamide-induced acute cardiotoxic rat models. Cardiovasc Toxicol. 2014; 14: 232–242.
[87]
Liepinsh E, Skapare E, Kuka J, Makrecka M, Cirule H, Vavers E, et al. Activated peroxisomal fatty acid metabolism improves cardiac recovery in ischemia-reperfusion. Naunyn-Schmiedeberg’s Archives of Pharmacology. 2013; 386: 541–550.
[88]
Lai YH, Chien Y, Kwok CF, Ho LT. Enhanced long-chain fatty acid uptake contributes to overaccumulation of triacylglycerol in hyperinsulinemic insulin-resistant 3T3-L1 adipocytes. Metabolism. 2010; 59: 1784–1793.
[89]
Li P, Liu Y, Zhang Y, Long M, Guo Y, Wang Z, et al. Effect of non-esterified fatty acids on fatty acid metabolism-related genes in calf hepatocytes cultured in vitro. Cellular Physiology and Biochemistry: International Journal of Experimental Cellular Physiology, Biochemistry, and Pharmacology. 2013; 32: 1509–1516.
[90]
Yan S, Yang XF, Liu HL, Fu N, Ouyang Y, Qing K. Long-chain acyl-CoA synthetase in fatty acid metabolism involved in liver and other diseases: an update. World Journal of Gastroenterology. 2015; 21: 3492–3498.
[91]
Jung HS, Shimizu-Albergine M, Shen X, Kramer F, Shao D, Vivekanandan-Giri A, et al. TNF-alpha induces acyl-CoA synthetase 3 to promote lipid droplet formation in human endothelial cells. Journal of Lipid Research. 2020; 61: 33–44.
[92]
Sandoval A, Fraisl P, Arias-Barrau E, Dirusso CC, Singer D, Sealls W, et al. Fatty acid transport and activation and the expression patterns of genes involved in fatty acid trafficking. Archives of Biochemistry and Biophysics. 2008; 477: 363–371.
[93]
Cruz-Gil S, Sanchez-Martinez R, Wagner-Reguero S, Stange D, Scholch S, Pape K, et al. A more physiological approach to lipid metabolism alterations in cancer: CRC-like organoids assessment. PLoS ONE. 2019; 14: e0219944.
[94]
Wang H, Airola MV, Reue K. How lipid droplets “TAG” along: Glycerolipid synthetic enzymes and lipid storage. Biochimica Et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 2017; 1862: 1131–1145.
[95]
Jiang YJ, Feingold KR. The expression and regulation of enzymes mediating the biosynthesis of triacylglycerols and phospholipids in keratinocytes/epidermis. Dermatoendocrinol. 2011; 3: 70–76.
[96]
Berry SE. Triacylglycerol structure and interesterification of palmitic and stearic acid-rich fats: an overview and implications for cardiovascular disease. Nutrition Research Reviews. 2009; 22: 3–17.
[97]
van Rooijen MA, Plat J, Blom WAM, Zock PL, Mensink RP. Dietary stearic acid and palmitic acid do not differently affect ABCA1-mediated cholesterol efflux capacity in healthy men and postmenopausal women: a randomized controlled trial. Clinical Nutrition. 2021; 40: 804–811.
[98]
Teh SS, Voon PT, Hock Ong AS, Choo YM. Incorporation of Palmitic Acid or Stearic Acid into Soybean Oils Using Enzymatic Interesterification. Journal of Oleo Science. 2016; 65: 797–802.
[99]
Khayyo VI, Hoffmann RM, Wang H, Bell JA, Burke JE, Reue K, et al. Crystal structure of a lipin/Pah phosphatidic acid phosphatase. Nature Communications. 2020; 11: 1309.
[100]
Gu W, Gao S, Wang H, Fleming KD, Hoffmann RM, Yang JW, et al. The middle lipin domain adopts a membrane-binding dimeric protein fold. Nature Communications. 2021; 12: 4718.
[101]
Bosma KJ, Kaiser CE, Kimple ME, Gannon M. Effects of Arachidonic Acid and Its Metabolites on Functional Beta-Cell Mass. Metabolites. 2022; 12: 342.
[102]
Nguyen P, Leray V, Diez M, Serisier S, Bloc’h JL, Siliart B, et al. Liver lipid metabolism. Journal of Animal Physiology and Animal Nutrition. 2008; 92: 272–283.
[103]
Xiong J. Fatty Acid Oxidation in Cell Fate Determination. Trends in Biochemical Sciences. 2018; 43: 854–857.
[104]
Adeva-Andany MM, Carneiro-Freire N, Seco-Filgueira M, Fernandez-Fernandez C, Mourino-Bayolo D. Mitochondrial beta-oxidation of saturated fatty acids in humans. Mitochondrion. 2019; 46: 73–90.
[105]
Console L, Giangregorio N, Indiveri C, Tonazzi A. Carnitine/acylcarnitine translocase and carnitine palmitoyltransferase 2 form a complex in the inner mitochondrial membrane. Molecular and cellular biochemistry. 2014; 394: 307–314.
[106]
Houten SM, Violante S, Ventura FV, Wanders RJ. The Biochemistry and Physiology of Mitochondrial Fatty Acid beta-Oxidation and Its Genetic Disorders. Annual Review of Physiology. 2016; 78: 23–44.
[107]
Talley JT, Mohiuddin SS. Biochemistry, Fatty Acid Oxidation. StatPearls. 2022. Available at: https://www.ncbi.nlm.nih.gov/books/NBK556002/ (Accessed: 24 January 2022).
[108]
Paton CM, Ntambi JM. Biochemical and physiological function of stearoyl-CoA desaturase. American Journal of Physiology. Endocrinology and Metabolism. 2009; 297: E28–E37.
[109]
Wang Z, Park HG, Wang DH, Kitano R, Kothapalli KSD, Brenna JT. Fatty acid desaturase 2 (FADS2) but not FADS1 desaturates branched chain and odd chain saturated fatty acids. Biochimica Et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 2020; 1865: 158572.
[110]
Park HG, Kothapalli KSD, Park WJ, DeAllie C, Liu L, Liang A, et al. Palmitic acid (16:0) competes with omega-6 linoleic and omega-3 a-linolenic acids for FADS2 mediated Delta6-desaturation. Biochimica et Biophysica Acta. 2016; 1861: 91–97.
[111]
Vriens K, Christen S, Parik S, Broekaert D, Yoshinaga K, Talebi A, et al. Evidence for an alternative fatty acid desaturation pathway increasing cancer plasticity. Nature. 2019; 566: 403–406.
[112]
Li YL, Tian H, Jiang J, Zhang Y, Qi XW. Multifaceted regulation and functions of fatty acid desaturase 2 in human cancers. American Journal of Cancer Research. 2020; 10: 4098–4111.
[113]
Rioux V, Lemarchal P, Legrand P. Myristic acid, unlike palmitic acid, is rapidly metabolized in cultured rat hepatocytes. The Journal of Nutritional Biochemistry. 2000; 11: 198–207.
[114]
Jakobsson A, Westerberg R, Jacobsson A. Fatty acid elongases in mammals: their regulation and roles in metabolism. Progress in Lipid Research. 2006; 45: 237–249.
[115]
Nie L, Pascoa TC, Pike ACW, Bushell SR, Quigley A, Ruda GF, et al. The structural basis of fatty acid elongation by the ELOVL elongases. Nature Structural & Molecular Biology. 2021; 28: 512–520.
[116]
Blacklock BJ. Fatty acid elongation by ELOVL condensing enzymes depends on a histidine nucleophile. Nature Structural & Molecular Biology. 2021; 28: 462–464.
[117]
Duca FA, Yue JT. Fatty acid sensing in the gut and the hypothalamus: in vivo and in vitro perspectives. Molecular and Cellular Endocrinology. 2014; 397: 23–33.
[118]
Efeyan A, Comb WC, Sabatini DM. Nutrient-sensing mechanisms and pathways. Nature. 2015; 517: 302–310.
[119]
Foster DA. Phosphatidic acid and lipid-sensing by mTOR. Trends in Endocrinology and Metabolism. 2013; 24: 272–278.
[120]
Menon D, Salloum D, Bernfeld E, Gorodetsky E, Akselrod A, Frias MA, et al. Lipid sensing by mTOR complexes via de novo synthesis of phosphatidic acid. The Journal of biological chemistry. 2017; 292: 6303–6311.
[121]
Saxton RA, Sabatini DM. mTOR Signaling in Growth, Metabolism, and Disease. Cell. 2017; 168: 960–976.
[122]
Gu X, Orozco JM, Saxton RA, Condon KJ, Liu GY, Krawczyk PA, et al. SAMTOR is an S-adenosylmethionine sensor for the mTORC1 pathway. Science. 2017; 358: 813–818.
[123]
Hoxhaj G, Hughes-Hallett J, Timson RC, Ilagan E, Yuan M, Asara JM, et al. The mTORC1 Signaling Network Senses Changes in Cellular Purine Nucleotide Levels. Cell Reports. 2017; 21: 1331–1346.
[124]
Zhao Y, Guo X, Yan S, Shi B, Sheng R. Acetate regulates milk fat synthesis through the mammalian target of rapamycin/eukaryotic initiation factor 4E signaling pathway in bovine mammary epithelial cells. Journal of Dairy Science. 2021; 104: 337–345.
[125]
Cheng J, Zhang Y, Ge Y, Li W, Cao Y, Qu Y, et al. Sodium butyrate promotes milk fat synthesis in bovine mammary epithelial cells via GPR41 and its downstream signalling pathways. Life sciences. 2020; 259: 118375.
[126]
He A, Chen X, Tan M, Chen Y, Lu D, Zhang X, et al. Acetyl-CoA Derived from Hepatic Peroxisomal beta-Oxidation Inhibits Autophagy and Promotes Steatosis via mTORC1 Activation. Molecular Cell. 2020; 79: 30–42.e4.
[127]
Li N, Zhao F, Wei C, Liang M, Zhang N, Wang C, et al. Function of SREBP1 in the milk fat synthesis of dairy cow mammary epithelial cells. International Journal of Molecular Sciences. 2014; 15: 16998–17013.
[128]
Mielenz M. Invited review: nutrient-sensing receptors for free fatty acids and hydroxycarboxylic acids in farm animals. Animal. 2017; 11: 1008–1016.
[129]
Julliard AK, Al Koborssy D, Fadool DA, Palouzier-Paulignan B. Nutrient Sensing: Another Chemosensitivity of the Olfactory System. Frontiers in Physiology. 2017; 8: 468.
[130]
Ma Y, Wang X, Yang H, Zhang X, Yang N. Involvement of CD36 in Modulating the Decrease of NPY and AgRP Induced by Acute Palmitic Acid Stimulation in N1E-115 Cells. Nutrients. 2017; 9: 626.
[131]
Liu H, Xu Y, Wang Y, Zhong S, Wang M, Lin P, et al. Cd36 is a candidate lipid sensor involved in the sensory detection of fatty acid in zebrafish. Physiology & Behavior. 2017; 182: 34–39.
[132]
Libran-Perez M, Polakof S, Lopez-Patino MA, Miguez JM, Soengas JL. Evidence of a metabolic fatty acid-sensing system in the hypothalamus and Brockmann bodies of rainbow trout: implications in food intake regulation. Integrative and Comparative Physiology. 2012; 302: R1340–R1350.
[133]
Martin C, Passilly-Degrace P, Chevrot M, Ancel D, Sparks SM, Drucker DJ, et al. Lipid-mediated release of GLP-1 by mouse taste buds from circumvallate papillae: putative involvement of GPR120 and impact on taste sensitivity. Journal of Lipid Research. 2012; 53: 2256–2265.
[134]
Dickson EJ, Falkenburger BH, Hille B. Quantitative properties and receptor reserve of the IP(3) and calcium branch of G(q)-coupled receptor signaling. Journal of General Physiology. 2013; 141: 521–535.
[135]
Den Hartogh DJ, Vlavcheski F, Giacca A, Tsiani E. Attenuation of Free Fatty Acid (FFA)-Induced Skeletal Muscle Cell Insulin Resistance by Resveratrol is Linked to Activation of AMPK and Inhibition of mTOR and p70 S6K. International Journal of Molecular Sciences. 2020; 21: 4900.
[136]
Kim JY, Lee HJ, Lee SJ, Jung YH, Yoo DY, Hwang IK, et al. Palmitic Acid-BSA enhances Amyloid-beta production through GPR40-mediated dual pathways in neuronal cells: Involvement of the Akt/mTOR/HIF-1alpha and Akt/NF-kappaB pathways. Scientific Reports. 2017; 7: 1–16.
[137]
Yasuda M, Tanaka Y, Kume S, Morita Y, Chin-Kanasaki M, Araki H, et al. Fatty acids are novel nutrient factors to regulate mTORC1 lysosomal localization and apoptosis in podocytes. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease. 2014 ; 1842: 1097–1108.
[138]
Matoba A, Matsuyama N, Shibata S, Masaki E, Emala CW, Sr., Mizuta K. The free fatty acid receptor 1 promotes airway smooth muscle cell proliferation through MEK/ERK and PI3K/Akt signaling pathways. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2018; 314: L333–L348.
[139]
Lager S, Jansson T, Powell TL. Differential regulation of placental amino acid transport by saturated and unsaturated fatty acids. American Journal of Physiology-Cell Physiology. 2014; 307: C738–C744.
[140]
Velasco C, Otero-Rodiño C, Comesaña S, Míguez JM, Soengas JL. Hypothalamic mechanisms linking fatty acid sensing and food intake regulation in rainbow trout. Journal of Molecular Endocrinology. 2017; 59: 377–390.
[141]
Chang YC, Lin CW, Chang YS, Chen PH, Li CY, Wu WC, et al. Monounsaturated oleic acid modulates autophagy flux and upregulates angiogenic factor production in human retinal pigment epithelial ARPE-19 cells. Life Sciences. 2020; 259: 118391.
[142]
Su Z, Zeng K, Feng B, Tang L, Sun C, Wang X, et al. Kun-Dan Decoction Ameliorates Insulin Resistance by Activating AMPK/mTOR-Mediated Autophagy in High-Fat Diet-Fed Rats. Frontiers in Pharmacology. 2021; 1301.
[143]
Shaw B, Lambert S, Wong MH, Ralston JC, Stryjecki C, Mutch DM. Individual saturated and monounsaturated fatty acids trigger distinct transcriptional networks in differentiated 3T3-L1 preadipocytes. Lifestyle Genomics. 2013; 6: 1–15.
[144]
Fang Y, Vilella-Bach M, Bachmann R, Flanigan A, Chen J. Phosphatidic Acid-Mediated Mitogenic Activation of mTOR Signaling. Science. 2001; 294: 1942–1945.
[145]
Hornberger TA, Chu WK, Mak YW, Hsiung JW, Huang SA, Chien S. The role of phospholipase D and phosphatidic acid in the mechanical activation of mTOR signaling in skeletal muscle. Proceedings of the National Academy of Sciences. 2006; 103: 4741–4746.
[146]
Cahill E, Salery M, Vanhoutte P, Caboche J. Convergence of dopamine and glutamate signaling onto striatal ERK activation in response to drugs of abuse. Frontiers in Pharmacology. 2014; 4: 172.
[147]
Franco R, Martinez-Pinilla E, Navarro G, Zamarbide M. Potential of GPCRs to modulate MAPK and mTOR pathways in Alzheimer’s disease. Progress in Neurobiology. 2017; 149: 21–38.
[148]
Alhosaini K, Azhar A, Alonazi A, Al-Zoghaibi F. GPCRs: The most promiscuous druggable receptor of the mankind. Saudi Pharmaceutical Journal. 2021; 29: 539–551.
[149]
Filippi-Chiela EC, Viegas MS, Thome MP, Buffon A, Wink MR, Lenz G. Modulation of Autophagy by Calcium Signalosome in Human Disease. Molecular Pharmacology. 2016; 90: 371–384.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top