Ceramides Are Fuel Gauges On The Drive To Cardiometabolic Disease (Wilkerson Et Al., 2024)
Ceramides Are Fuel Gauges On The Drive To Cardiometabolic Disease (Wilkerson Et Al., 2024)
2
3 Joseph L. Wilkerson, Sean M. Tatum, William L. Holland, and Scott A. Summers†
4 Department of Nutrition and Integrative Physiology, University of Utah, Salt Lake City, UT 84112
5
6
7 Department of Nutrition and Integrative Physiology, University of Utah College of Health
8
9 Ṫ Correspondence should be addressed to:
10 Scott A. Summers
11 Department of Integrative Physiology
12 University of Utah College of Health
13 15N, 2030 East, Rm 3110
14 Salt Lake City Utah 84112
15 Email: [email protected]
16 Tel: 801-585-9359
17
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
21
22 ABSTRACT
23 Ceramides are signals of fatty acid excess that accumulate when a cell’s energetic
24 needs have been met and its nutrient storage has reached capacity. As these sphingolipids
25 accrue, they alter the metabolism and survival of cells throughout the body including in the
26 heart, liver, blood vessels, skeletal muscle, brain, and kidney. These ceramide actions elicit the
27 tissue dysfunction that underlies cardiometabolic diseases such as diabetes, coronary artery
29 biosynthesis and degradation pathways which maintain ceramide levels in normal physiology
30 and discuss how the loss of ceramide homeostasis drives cardiometabolic pathologies. We
31 highlight signaling nodes that sense small changes in ceramides, and in turn reprogram cellular
32 metabolism and stimulate apoptosis. Lastly, we evaluate the emerging therapeutic utility of
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
33 these unique lipids as biomarkers that forecast disease risk and as targets of ceramide-lowering
• Sphingolipids such as ceramides are signals of fatty acid excess that increase
disorders.
35
36 INTRODUCTION
38 steatohepatitis (MASH), and heart failure are afflicting the population at unsustainable rates,
39 driven by decreased physical activity and increased availability of diverse, highly palatable,
40 calorie-rich foods (1). Most excess calories are stored in fatty acids contained within
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
42 phosphatidylcholine, phosphatidylethanolamine, etc.) present in lipid bilayers. A smaller
43 proportion of fatty acids resides in membrane-bound sphingolipids, which are produced when
44 they attach to an amino acid-based scaffold rather than glycerol. Though sphingolipids were
48 health outcomes and causal factors that contribute to the tissue dysfunction that underlies
50 of ceramide synthesis and degradation. These regulatory events are being discovered at a
51 growing pace, owing to technical innovations that allow for precise quantification of steady-state
52 sphingolipid levels and discernment of dynamic changes in ceramide biosynthesis and turnover.
53 In individuals with the metabolic syndrome, various nutritional, hormonal and inflammatory
54 signals disrupt these homeostatic processes to promote accumulation of ceramides, which act
55 as cellular signals of fatty acid excess and fundamental drivers of cellular stress (2-9). Indeed,
56 serum ceramides correlate with insulin resistance, coronary artery disease, major adverse
59 including insulin resistance, hypertension, and atherosclerosis, as well as hepatic, cardiac, and
60 vascular dysfunction (2, 4, 11-37). Because of these observations, researchers are now
61 exploring the clinical utility of diagnostics and therapeutics that target these unique molecules
62 (38, 39). Herein we review these remarkable findings and discuss the still unresolved
66 physician J.L.W. Thudichum described a surprising discovery about the organic components of
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
67 the brain. He identified a peculiar form of lipid that did not contain glycerol, the main scaffold of
68 most phospholipids, and instead comprised an unknown moiety he termed sphingosine (40).
69 The most abundant of these unknown lipid molecules was sphingomyelin, which contained a
70 fatty acid, phosphoric acid, and two bases: choline and sphingosine. Thudichum added the
71 prefix “sphingo” to the curious lipid forms to denote their enigmatic nature. Taken from the
72 Sphinx in Greek mythology who confounded travelers along to road to Thebes with her riddles,
73 the term served as a reminder that these lipids were likely to be doing things no one at the time
74 could comprehend (40). Nonetheless, in the years immediately following Thudicum’s discovery,
75 this novel and distinct lipid class was largely ignored (41).
77 The structures of most sphingolipids were not resolved until the 1940s, and an
78 understanding of their biological functions was even slower to materialize. Initially, these minor
79 constituents of cell membranes were relegated to the role of structural building blocks of
80 bilayers. This view changed in 1986 when Robert Bell and colleagues determined that
81 sphingosine participated in cell signaling: they determined that it was a competitive inhibitor of
82 protein kinase C that tempered enzyme activation by calcium and sn-1,2-dioleoylglycerol (42-
83 44). Beyond showing that sphingosine could inhibit protein kinase C in vitro, this collaborative
84 group of investigators demonstrated that sphingosine inhibited the enzyme in platelets (42),
85 neutrophils (43), and human promyelocytic leukemia cells (44). In neutrophils, this sphingosine
86 action blocked PKC activation of NADPH oxidase (43) to modulate the oxidative burst—
88 aids in killing invading bacteria (45-47). Concurrently, the team determined that sphinganine
89 and other long-chain sphingoid bases (e.g., long carbon chain sphingosines) inhibited PKC to
90 alter differentiation of human promyelocytic leukemia cells (44). These promyelocytic cells
91 mature into various other cell types, such as macrophages and monocyte-like cells, depending
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
92 on extrinsic stimulatory factors (48-50). The researchers found that the addition of long-chain
93 sphingoid bases to these cells inhibited PKC, leading to a loss of adhesion and differentiation
94 (44).
95 This series of papers implicating sphingosine in cell signaling was initially met with
96 resistance. Few believed that sphingosine and sphinganine, largely considered lipid byproducts
97 produced during the breakdown of more abundant and complex sphingolipids, played a
98 quantitatively important role as an intracellular messenger (40). However, the data piqued
99 curiosity and research interest in a possible signaling role for sphingolipids started to grow.
101 Shortly after sphingosine was found to be a signaling entity, researchers discovered that
103 In the early 1990s, Donald Gill and Sarah Spiegel found that S1P could elicit an entirely new
104 spectrum of actions, with Gill determining that it triggered calcium release from intracellular
105 stores (51) and Spiegel reporting that it could induce cellular proliferation (52). Shortly
106 thereafter, Timothy Hla discovered that S1P serves as a ligand for a G-protein coupled receptor
107 that influences endothelial cell function (53). More recently Hla and colleageus found that ApoM
108 serves as an S1P chaperone that carries it in HDLs and presents it to the receptor (54-56). We
109 now know that extracellular S1P serves as a ligand for at least five different receptors that are
110 implicated in a wide range of physiological process (57-62). Since those early discoveries, the
111 kinases that phosphorylate sphingosine and the G-protein coupled receptors that mediate S1P
112 actions were cloned and characterized (63). Moreover, therapeutics targeting the S1P receptors
113 are now in clinical use (63). For a more complete summary of the fascinating and diverse
114 subject of S1P biology, we refer the reviewer to several outstanding recent reviews (57, 59-61,
115 64-72).
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
116 Ceramides Emerge as Novel ‘Second Messengers’ that Link Inflammatory Cytokines to
117 Apoptosis
118 As sphingosine and S1P were emerging as signaling molecules, the Hannun and
120 coupled to an esterified free fatty acid—were stress-induced lipids that drove cell death. The
121 work stemmed from their discovery of a sphingomyelin cycle that involved acute liberation of
122 ceramide from sphingomyelin by sphingomyelinases (73-76). Excitingly, these research groups
123 separately determined that the inflammatory cytokine tumor necrosis factor-alpha (TNFα)
124 dramatically upregulated sphingomyelin turnover (77), prompting their speculation that ceramide
125 was a second messenger that would mediate cellular responses to inflammation and stress (78,
126 79). Kolesnick’s group determined that the TNFα dramatically increased ceramide and
127 decreased sphingomyelin levels in a cell free system, confirming the tight coupling of the TNFα
128 receptor to sphingomyelinase activity. Shortly thereafter, Obeid and Hannun discovered that
129 ceramides were potent drivers of apoptosis (8), shepherding in a new era of research on the
131 colleagues followed with the discovery that ceramides produced from the de novo synthesis
132 pathway, rather than via sphingomyelinases, also induced cell death (80), and that human
133 lymphoblasts and mice lacking sphingomyelinase were protected from radiation-induced
134 apoptosis (81). Controversy soon erupted, as skeptics argued that ceramides did not increase in
135 concordance with the apoptosis timeline or that ceramide analogs were eliciting a non-apoptotic
136 and non-specific form of cell death (82, 83). However, the subsequent availability of tools to
137 more precisely measure or manipulate ceramides soon resolved the transient debate, and
138 ceramides are now well-established drivers of apoptosis in a broad range of cell types and
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
141 Ceramides are the central building block of most complex sphingolipids, and their
142 cellular levels are controlled by pathways that influence their de novo synthesis (Figure 1) or
145 Palmitoyl-CoA and Serine Condense to Form the Canonical Sphingoid Base
146 De novo sphingolipid biosynthesis begins with the formation of sphinganine in the
147 endoplasmic reticulum (Figure 1). Serine palmitoyltransferase (SPT) catalyzes the initial
148 reaction (84), condensing an amino acid and an acyl-CoA to create the basic sphingoid
149 structure (85). The SPT complex comprises SPTLC1, 2, or 3 subunits that form heterodimers
150 with varying CoA preferences; SPTLC1 and 2 heterodimers show preference for the most
151 common substrates palmitoyl-CoA and serine, which condense to create the canonical 18-
152 carbon sphingoid scaffold found in most sphingolipids. This product is 3-ketosphinganine, which
154 include SPTLC1 and 3 show greater promiscuity for amino acids and fatty-acyl CoAs, which can
155 lead to the formation of less abundant but potentially important sphingoid moieties (86-88).
156 Additional regulatory subunits named SSTPSA and SSTPSB further modulate SPT activity and
158 The SPT complex can occasionally incorporate alanine and glycine, instead of serine, to
159 create a sphingoid base lacking the easily modifiable hydroxyl group at the C1 position (90, 91).
161 has been observed in individuals with hereditary sensory neuropathy (HSN)(92); patients with
162 HSN have gene variants in SPT subunits that alter the enzyme’s substrate preference, such that
163 they incorporate alanine or glycine more readily (92). Deoxysphingolipids are known to damage
164 neurons and potentially other cell types, and their accumulation could contribute to the tissue
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
166 SPT can also use fatty acid substrates other than palmitoyl-CoA to create sphingoid
167 bases of varying carbon chain lengths. A small number of studies suggest distinct roles for
168 these atypical and far less abundant sphingolipids (103, 104), but they are poorly understood.
170 The next enzymes in the de novo synthesis pathway, sphinganine N-acyltransferases,
171 add an acyl-chain to the amino group of the sphingoid bases (Figure 1). These acyltransferases
172 are typically termed ceramide synthases (CERS), even though they produce dihydroceramides
173 during canonical de novo sphingolipid biosynthesis (87, 105, 106). The CERS family of
174 enzymes can add multiple different N-acyl chains to the sphingoid moiety, and are thus
175 responsible for the tremendous variation seen in sphingolipids. Indeed, the sphingoid base can
177 The Futerman laboratory has extensively characterized this enzyme family, finding that
178 mammals have six ceramide synthases (CERS1-6) with differing tissue distributions and distinct
179 acyl-CoA specificities (105, 107): CERS1 adds C18-containing fatty acids to the sphingoid base
180 (108); CERS2 adds very long chain C22 to C24 fatty acids (109); CERS3 adds very long chain
181 fatty acids that are C26 or longer (110); CERS4 adds fatty acid chain lengths C18-C22 (111);
182 CERS5 and CERS6 generally add shorter chain fatty acids, mainly C16, though CERS6 is also
183 known to also add C14 onto the base (111, 112). The CERS enzymes generally show
184 preference for saturated fatty acyl-CoAs, though the CERS2 enzyme also transfers the
186
187 Though the CERS enzymes show specificity for specific fatty acid chains, they are
188 promiscuous with regards to the sphingoid backbone: they can N-acylate any of the sphingoid
189 bases described above, as well as synthetic mimics like fumonisin B1, NBD-sphinganine, and
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 1: De novo synthesis of ceramides. The
synthesis of new ceramides from palmitoyl-CoA and
the amino acid serine begins with the serine
palmitoyltransferase enzymatic complex. Acyl
chains, R, are inserted into the lipids via the CERS1-
6 enzymes, which have specificity for fatty acids with
defined carbon chain lengths. Next, a double bond is
formed in the sphingoid backbone by DES1 or
DES2. Common pharmaceutical inhibitors that
inhibit each step are shown in red.
190 fingolimod (113-115). The presence of such a large family of similar yet specialized CERS
191 enzymes presents many interesting questions (116). How are CERS enzymes regulated? Why
192 would cells show unique sphingolipid acyl-chain patterns, with tissues like the liver enriched in
193 very long chain (C24 and C24:1) species, while muscle is predominantly long-chain C18
10
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
194 species? How do sphingolipids with dissimilar chain lengths differ biochemically and
195 functionally? Later in this review we will discuss the body of literature suggesting that the C16-
196 containing ceramides are more pathogenic than very long chain ceramides in a variety of
198 A substantial amount of data suggest that CERS enzymes work in concert, and that
199 changes in one isoform will lead to compensatory alterations in others (109, 117-120). This
200 compensatory response may be possible because of the ability of CERS enzymes to form
201 heterodimers (119). In HeLa cells CERS2, 5, and 6, will form heterodimers after radiation
202 treatment (121). Using human embryonic kidney and hepatocellular carcinoma cell lines,
203 Futerman and colleagues demonstrated that CERS2 activity can be influenced by CERS5 and
204 CERS6 (119). These findings suggest that allosteric mechanisms control the activity of the
205 CERS family of proteins. Various post-translational modifications can also affect their activity
207 Futerman’s lab also found that S1P modulates CERS2 by acting as a non-competitive
208 inhibitor. This action pertains only to CERS2 and is possible because a small region of the
210 This unique domain is not present on the other CERS proteins. S1P binds to this region to effect
211 a conformation change that inhibits CERS2 activity (109). This finding may explain a
212 longstanding observation that increasing intracellular S1P by either upregulating sphingosine
213 kinases or downregulating an S1P phosphatase reduces levels of ceramides (123-126). The
215 interaction with other proteins, or feedback loops warrants further attention.
216 Sociale and colleagues reported that CERS enzymes could also regulate cell function
217 via mechanisms that were independent of their acyltransferase activity (127). They determined
218 that the proteins served as transcriptional repressors that modulated expression of lipase
11
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
219 genes. The observation stemmed from prior observations that mammalian (i.e., CERS1,2,4,5
220 and 6) and the Drosophila CERS homologue schlank contain nuclear localization sequences
221 and homeobox-like domains. The authors utilized knock-in constructs to discern the function of
222 schlank domains, finding that the flow of fatty acids through the sphingolipid synthesis pathway
223 elicited the translocation of schlank to the nucleus, where it suppresses transcription of the
224 lipase genes lipase 3 and magro (127). This work suggested the presence of an alternative
225 mechanism of CERS action. Nonetheless, some controversy exists. The HOX domains in the
226 CERS enzymes display irregularities as compared to other HOX-domains, as they were devoid
227 of residues considered essential for DNA binding (128, 129). Additional studies are necessary to
229 An additional acylation can occur on the free hydroxyl group of the serine backbone to
230 generate 1-O-acyl ceramides, which is a triglyceride-like form of sphingolipid that is likely
231 packaged into lipid droplets as a storage pool of sphingolipid (130). The canonical diglyceride
232 synthesizing enzyme diacylglycerol acyltransferase 2 (DGAT2) can also esterify ceramide.
233 Interestingly, acyl-CoA synthetase 5 directly interacts with ceramide synthases and DGAT2 to
234 divert fatty acids onto ceramides. Blocking the O-acyl esterification of ceramide increases
236 preventing the metabolism of ceramide into glucosylceramide increases levels of 1-O-
237 acylceramide, suggesting that the glucosylation or acylation of ceramide may be competing
238 pathways (131). Strategies that lower hepatic steatosis also decrease 1-O-acylceramides
239 (132). Of note, a 1-O-acylceramide synthase that transfers an acyl group from
240 phosphatidylethanolamine to the free hydroxyl group of ceramide has also been described
241 (133).
242 A Desaturase Inserts a Critical Double Bond into the Sphingoid Base
12
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
243 The final reaction in the de novo ceramide synthesis cascade inserts a key double bond
244 into dihydroceramides to produce the more abundant ceramides (84). Two distinct
245 dihydroceramide desaturases (DES1 and 2) catalyze this reaction, with DES1 serving as the
246 predominant enzyme isoform in most mammalian cell types (134, 135). The less abundant
247 DES2 enzyme is prevalent in skin, intestine, and kidney. DES2 also contains c-4 hydroxylase
248 activity and can generate phytoceramides, in addition to producing dihydroceramides (134).
249 Distinct Transport Machinery Moves Ceramides to the Golgi to Generate Complex
250 Sphingolipids
251 Ceramides are the first sphingolipid in the de novo biosynthesis pathway that accrues to
253 sphingolipids, these lipids serve as signals of increased pathway flux that alter the metabolic
255 After being formed in the endoplasmic reticulum, ceramides traffic to the Golgi apparatus
256 for further modification. Addition of various headgroups to the C1-hydroxy position can create a
257 multitude of different lipid types (Figure 2). For example, the addition of phosphorylcholine, via
258 sphingomyelin synthase, creates sphingomyelin, the most abundant sphingolipid and an
259 important binding partner of cholesterol in plasma membranes (137). Alternatively, the addition
260 of sugars including lactose, galactose, and glucose creates the complex glycosphingolipids and
262 The transport of ceramides between the ER to the Golgi involves at least one Ceramide
263 Transfer Protein (CERT), which preferentially delivers ceramides to sphingomyelin synthases
264 (139-141). CERT is a splice variant of the goodpasture antigen-binding protein (GPBP) that was
265 originally discovered in patients with the autoimmune Goodpasture syndrome (142). Following
266 its export from cells, GPBP phosphorylates collagen IV, which is recognized by antibodies that
13
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
267 accumulate to disrupt basement membranes. GPBP is primarily expressed in neural tissue
268 during development but is upregulated in autoimmune diseases (143, 144). The CERT splice
269 variant lacks a serine-rich 26 amino acid segment of the full GPBP protein (139). Unlike GPBP,
270 CERT is expressed in all cell types and acts as a ceramide carrier that can shuttle ceramides
271 throughout the cell (142). In mice, knocking out the Gpbp gene or introducing inactivating CERT
272 mutations produces embryonic death owing to destabilization of the endoplasmic reticulum and
274 CERT prefferentially transports ceramides with fatty acid chain lengths between 14 and
275 20 carbons, though it can also traffic longer carbon chain ceramides with lower affinity (139,
276 146, 147). This transport to the cis-Golgi is essential for the formation of sphingomyelin by
278 ceramide. Sphingomyelin is then transported by vesicular mechanisms through the Golgi and
279 other cellular compartments before finally be incorporated into cellular membranes (137).
280 Expression of the CERT proteins decreases during stress and increases in certain cancers and
281 infectious disease (148-151), leading to changes in the relative abundance of ceramide vs.
282 sphingomyelin in each of these conditions. For example, the decrease in CERT during stress
283 leads to elevations in ceramides and subsequent induction of apoptosis (148). CERT is also
284 regulated by protein kinase D and is ATP dependent (152, 153). Little is known about the
286 CERT may also shepherd sphingomyelin through the Golgi stacks and even possibly to
287 other membranes (106). It also removes ceramides from the endoplasmic reticulum membrane
288 or the plasma membrane. This activity depends upon the arrangement of ceramides within
289 membranes. For instance, CERT struggles to remove ceramides that are in tightly packed
14
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
291 Ceramides can also be phosphorylated by a ceramide kinase to generate ceramide-1-
292 phosphate (C1P, Figure 2)(156). Human C1P transfer protein (CPTP) selectively transports
293 C1P from the Golgi apparatus to other cellular subdomains through a non-vesicular trafficking
294 mechanism (157). C1P has various physiological effects in a cell. It has been shown to mediate
295 cell survival and decrease apoptosis, increase iNOS, and promote inflammatory responses.
296 Many of these actions are done through regulation of the SPT complex and acidic
297 sphingomyelinase (157). In cultured macrophages, C1P can regulate glucose uptake by
298 promoting the translocation of GLUT3 to the plasma membrane through PIP3K/AKT signaling
299 (158).
300 Ceramides also traffic via vesicular mechanisms to the cytosolic leaflet of the cis-Golgi
301 sub-compartments that produce glucosylceramide (159, 160). Four-phosphate adaptor protein 2
302 (FAPP2) transports glucosylceramide to the trans-Golgi, where it receives the additional
303 carbohydrate moieties to become complex glycosphingolipids and gangliosides (161). Similarly,
304 sulfatide can be added to galactosylceramide to produce sulfatide, a critical lipid in the myelin
305 sheath that coats neurons. These glucosylceramide derivatives play key roles in cellular
306 recognition, as they display remarkable diversity owing to thre presence of a variety of acyl
307 chains, as well as head groups comprising glucose, galactose, galactosamine, sialic acid, etc.
308 (Figure 2). For example, GM3 synthase adds sialic acid to lactosylceramide to generate GM3
309 ganglioside as the first step in the synthesis of more complex gangliosides. The addition of such
310 a diverse array of carbohydrate residues ultimately produces more than 500 glycosphingolipid
15
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
331
16
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
332 Complex Sphingolipids are Broken Down in Lysosomes
334 low pH—liberate ceramides from complex sphingolipids such as sphingomyelin and
336 an acidic acid ceramidase (ASAH1) to produce sphingosine and a free fatty acid. The
337 sphingosine is capable of being re-acylated by CERS enzymes to recreate ceramides via the
340 To date, nine distinct inborn errors of metabolism, termed sphingolipidosis, have been
341 identified. These disorders result from an impaired ability to degrade glucosylceramide
342 derivatives within the lysosome. Tay-Sachs and Sandhoff diseases over accumulate GM2
344 common feature for most of these lysosomal storage diseases, likely owing to the lack of self-
346 impairments in the biosynthesis of gangliosides, as seen with GM3 synthase deficiency, impair
347 neuronal development and trigger epilepsy. Although gangliosides have been commonly
348 studied in the context of these rare diseases, they have also emerged as potential mediators of
352 (170). Six different genes encode the sphingomyelinase family members, which include
17
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
356 SMPD1, commonly called acid sphingomyelinase, displays highest activity in the acidic
357 environment of the lysosome, while also retaining suboptimal but biologically relevant activity at
358 a neutral pH (165, 172, 173). The precursor SMPD1 protein contains six potential glycosylation
359 sites that modify protein function, protect it from proteolysis, direct it to specific locations within
360 the cell, or label it for secretion (174, 175). The addition of a mannose targets the enzyme to
361 lysosomes, where it can operate in its preferential environment to hydrolyze sphingomyelin.
362 Acid sphingomyelinase lacking this mannose moiety can function at a neutral pH, but requires
363 zinc as a cofactor (174, 176). In this form, SMPD1 can act in caveolae at the plasma membrane
364 (177) or be secreted from the cell, where it can function in the extracellular space (172, 178).
365 The SMPD2-5 genes encode neutral sphingomyelinases, which are commonly termed n-
366 SMase1, n-SMase2, n-SMase3, and MA-SMase, respectively. They require a Mg2+ ion for
367 catalytic activity (179). The ubiquitous n-SMase2 is best characterized; it works at the plasma
368 membrane and is present in all tissues, with highest expression in the brain. The protein is
369 palmitoylated in two locations, and these acyl-chains anchor it to the plasma membrane (180,
370 181). n-SMase1 is a much smaller protein than n-SMase2, comprising only 423 amino acids,
371 compared to 655 amino acids in n-SMase2. This protein mainly localizes to the endoplasmic
372 reticulum, but it has also been found in the nucleus and Golgi (180, 182, 183). n-SMase3 is the
373 largest protein of the family (866 amino acids) and shares no homology with the other neutral
374 sphingomyelinases. It is highly abundant in cardiac muscle tissue, where it resides in the
375 endoplasmic reticulum and the trans-Golgi network (184). The final neutral sphingomyelinase is
377 homology and structural similarities to n-SMase2. MA-SMase is found in many tissues including
378 the brain, testis, and skin. The protein preferentially migrates to mitochondria, but it has also
18
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
380 The third type of sphingomyelinase works at an alkaline pH. Alkaline sphingomyelinase
382 does not share sequence or structural homology with any of the other sphingomyelinases and
383 does not require cofactors to be active; in fact, zinc ions inhibit catalytic activity. It shows
384 preferential expression in the intestine where it may hydrolyze dietary sphingomyelin (187, 188).
386 Ceramidase enzymes excise the acyl-chain from ceramide to acutely split it into its two
387 component parts: a sphingoid base and a free fatty acid (189). Mammals were originally thought
388 to express five ceramidases which were classified by their pH optima: one acid ceramidase
389 (ASAH1); one neutral ceramidase (ASAH2); and three alkaline ceramidases (ACER1-3). More
390 recently, a family of progestin and adipoQ (PAQR) superfamily receptors—which includes the
391 adiponectin receptors—were found to be lipid hydrolases that deacylate ceramides (2, 190).
392 This discovery enlarged the family of ceramidase enzymes capable of acutely lowering
395 environments, was the first ceramidase to be characterized. Gatt and colleagues discovered the
396 enzyme in 1966 from rat brain extracts (191). It has been extensively studied, in part, because
397 of a 1972 discovery by Sugita et al. that inactivating ASAH1 mutations caused Farber’s disease,
398 a rare and fatal lysosomal storage disorder (192). ASAH1 is ubiquitously expressed, with very
399 high expression in the heart and kidneys. The enzyme functions at optimal catalytic activity at a
400 pH of 4.5 and operates in lysosomes (193), though it has been implicated in non-lysosomal
401 functions as well. It displays a preference for ceramides with unsaturated 6 to 16 carbon chain
402 fatty acids (194). Interestingly, ASAH1 also catalyzes the reverse reaction; Okino et al.
403 demonstrated that addition of sphingosine and a C12:0 fatty acid to the enzyme in vitro leads to
404 the reformation of ceramides. The reverse reaction is optimal at a higher pH (pH = 6)(195),
19
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
405 suggesting that it likely operates at non-lysosomal cellular locales. Germline depletion of the
406 Asah1 gene is lethal for mice, as embryos fail to progress past the 2-cell stage (196). However,
407 conditional knockout of Asah1 has proven to be a useful tool for inducing ceramides in adult
408 mouse tissues, and these models have shed important insight about the functional
410 Other ceramidases have received less attention. Neutral ceramidase (ASAH2) is
411 expressed highly in the liver, small intestine, and colon where it is thought to contribute to the
412 digestion of dietary sphingolipids (198). The enzyme has also been identified in other tissues
413 (e.g., kidney), suggesting that it has additional roles that are unrelated to dietary absorption
414 (199-202). The alkaline ceramidases, ACERs 1-3, function at pH ~8. ACER1 is highly
415 expressed in the skin, particularly in epidermal keratinocytes. It is localized to the endoplasmic
416 reticulum and is required to maintain skin homeostasis (203). It preferentially hydrolyzes
417 substrate ceramides that contain an unsaturated very long chain fatty acid (204), which are
418 enriched in the epidemis. ACER2 localizes to the Golgi and is highly expressed in the placenta,
419 stomach, and bladder, particularly in cells that secrete the gastric mucosa. This calcium-
420 dependent enzyme is involved in apoptosis and necrosis of cells after DNA damage (205-207).
421 ACER3, discovered in 2001, was the first alkaline ceramidase found in mammals. It localizes to
422 both the endoplasmic reticulum and the Golgi, and is ubiquitously expressed, with particularly
423 very high levels in the placenta. ACER3 catalyzes the hydrolysis of ceramides containing an
424 unsaturated long fatty acid chain. The enzyme is essential for cell proliferation and regulating
426 After ceramidases hydrolyze ceramide, the sphingoid base can either be recycled into
427 other ceramides via the salvage pathway or it can be phosphorylated by the sphingosine
428 kinases to make sphingosine-1 phosphate (S1P). In the salvage pathway, released sphingoid
20
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
430 Production of Sphingosine-1-Phosphate and Subsequent Degradation of Sphingolipids
431 The sphingosine kinases that generate S1P are products of two distinct genes: SPHK1
432 and SPHK2. The enzymes are largely homologous, with the notable exception that SPHK2
433 selectively possesses nuclear localization and nuclear export signals, as well as a lipid binding
434 domain at the N-terminal (211-213). Humans and mice contain two SPHK1 and three SPHK2
435 enzymes, each being a product of alternative gene splicing (214). The kinases generate
436 different pools of S1P in discrete subcellular regions that have differential effects on cellular
437 behavior (215). S1P can then be exported from the cell via a spinster protein (SPNS2) (216),
438 where it signals via the hormonal, autocrine or paracrine mechanisms discussed earlier (217).
440 in the endoplasmic reticulum (218, 219). These reactions regenerate sphingosine, which can be
441 used to re-build ceramides via the aforementioned salvage pathway. Ultimately, destruction of
442 the sphingoid base involves the enzyme S1P lyase, which converts S1P to ethanolamine
443 phosphate and hexadecanal (220). This reaction eliminates the sphingoid structure and permits
444 permanent exit of the component building blocks from the sphingolipid pathway (63).
446 Palmitate and serine are critical for the committed steps in sphingolipid biosynthesis, and
447 their availability undoubtedly influences rates of ceramide production. Inflammatory agonists,
448 adipokines, and the microbiome also influence rates of ceramide synthesis and degradation (4,
449 28). Since little is known about the post-translational and allosteric regulation of sphingolipid
450 biosynthetic enzymes, transcriptional changes in enzymatic machinery have served as key but
451 often imprecise readouts of enzyme activity. With advances in mass spectrometry-based
452 approaches for measuring sphingolipid flux, we anticipate that our understanding of regulatory
453 mechanisms governing sphingolipid biosynthesis and degradation will soon grow.
21
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
454 Regulatory Interacting Proteins Influence the Activity of Key Biosynthetic Enzymes
455 The tight control of cellular ceramide levels suggests the existence of feedback and
456 other mechanisms for titrating sphingolipid abundance (221). The Weissman lab determined
457 that ORM proteins in yeast and ORMDL proteins in humans inhibit SPT activity to finely tune
458 rates of ceramide biosynthesis (222). The function of these ER-localized transmembrane
459 proteins was first discerned using functional genomic screens of yeast libraries, which revealed
460 reciprocal expression patterns between certain ORM transcripts and SPT subunits. These data
461 suggested that these proteins had opposing cellular functions. Indeed, genetic overexpression
462 of ORM proteins reduced ceramides and the Lcb transcripts that encode SPT subunits.
463 Moreover, in human cells, genetic knockdown of all 3 uniquely coded ORMDL isoforms
464 increased ceramides ~3-fold. SPT and ORMDL proteins physically interact with one another,
465 which allows the ORMDL proteins to blunt SPT’s catalytic activity. Studies with myriocin
466 confirmed that the ORM proteins were feedback regulators of sphingolipid synthesis, as the
467 compound relieved the functional brakes on SPT by stimulating their phosphorylation. Follow-
468 up studies of this regulatory phosphorylation event implicated yeast protein kinase 1 (Ypk1) and
469 the morphogenesis checkpoint kinase Swe1 as candidate ORM kinases (223-225). These
470 enyzmes are activated in yeast by myriocin or heat stress (223-225). Curiously, mammalian
472 Sasset and colleagues found that S1P stabilizes ORMDL3 to increase its inhibitory
473 activity and slow sphingolipid biosynthesis in melanoma cell lines(227). Noting that ORMDL
474 sequences contain a conserved prolyl hydroxylase (PHD) consensus domain, known to regulate
475 the levels of hypoxia-inducible factor 1α (HIF-1α)(228) and other proteins (229) via
476 ubiquitination, they tested whether ER-membrane associated degradation of ORMDL served as
477 a regulatory mechanism to control their stability. Their data showed that PHD-mediated
478 hydroxylation of Pro137 enforces constitutive degradation of ORMDLs, which have a relatively
22
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
479 short half-life. Using multiple genetic and pharmacological strategies they demonstrated that
480 disrupting S1P sensing enhances flux through the de novo ceramide biosynthetic pathway. S1P
481 signaling inhibited PHD-mediated Pro137 hydroxylation to stabilize ORMDLs and downregulate
482 SPT activity. These studies supported previous observations that sphingosine kinases are
483 negative regulators of SPT (230). ER ceramides may trigger a similar feedback mechanism, as
484 they were shown to blunt SPT activity in cell-free isolated membrane preps containing ORMDL
486 Impressive genetic studies by the Dunn and Bönnemann laboratories shed further light
487 on the mechanisms and relevance of the interaction of ORMDLs with SPT (232). They
488 identified four distinct genetic coding variants within SPTLC1 in patients with childhood-onset
489 amyotrophic lateral sclerosis, finding point mutations A20S and Y23F, as well as deletion
490 variants Δ39 and Δ40,41. These variants, which were identified in seven independent families,
491 were in highly conserved residues residing near the transmembrane domain. The sites are
492 critical for the binding of ORMDL proteins, and their variants allow the SPTLC subunits to
493 operate without restraint. Genetic dosing of ORMDL3 during overexpression studies
494 demonstrated that these SPTLC1 mutants were refractory to ORM-mediated inhibition in SPT
495 activity (232). Patients with these mutations develop amyotrophic lateral sclerosis (ALS), a
497 Beautiful work by the Di Lorenzo lab defined NOGO-A and NOGO-B proteins as another
498 class of molecules that inhibit SPT activity (221, 233-235). Three splice variants of Reticulon-4
499 (Rtn4), the gene encoding NOGO proteins, , have been described. NOGO-B is highly produced
500 in vascular endothelium, and its genetic deletion upregulates SPT activity by ~40% (235).
501 Similarly, genetic ablation of Rtn4A, encoding NOGO-A, which is transiently expressed in
503 pressure overload (236). A third NOGO variant, NOGO-C which is enriched in the central
23
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
504 nervous system, may play a role in cancer biology, and its knockdown blunts Akt-
505 phosphorylation, an effect consistent with elevated ceramides (237). Although NOGO co-
506 precipitates with SPT and ORMDL complexes, the precise molecular regions that control
509 The availability of palmitate and serine influences rates of ceramide synthesis (238) and
510 sphingolipids are known to accrue in mice and rats fed obesogenic diets enriched in saturated
511 fats (13, 239)(Figure 3). We have shown that intravenous infusion of lipid emulsions enriched in
512 saturated fats, but not those that are predominantly unsaturated, into rats drives ceramide
24
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
513 accumulation in multiple tissues (4, 13). A small number of dietary studies have shown that fat
514 consumption influences ceramide levels in humans, with diets high in saturated fat increasing
515 ceramide levels and worsening metabolic outcomes (240, 241). Indeed, Luukkonen et al.,
516 demonstrated that humans that were overfed saturated fats for three weeks showed
517 accumulation of harmful C16-containing ceramides in the liver which was associated with the
518 development of insulin resistance (240). Similarly, Muoio and colleagues found that feeding
519 humans diets that were high in palmitate led to increased levels of muscle and serum
520 ceramides, while diets that were high in oleate did not (241). Bariatric surgery rapidly
522 Metallo and colleagues recently explored the influence of amino acid bioavailability on
523 sphingolipid biosynthesis (99, 246). Complete removal of serine from the diet lowers ceramide
524 levels in mice, leading to compensatory elevations in deoxysphingolipids derived from either
525 alanine or glycine (246). His group reported that mouse models of type 1 and type 2 diabetes
526 show low systemic serine and glycine concentrations in liver and serum, as well as increased
527 serine clearance during a serine tolerance test (99). This finding is consistent with the
528 enhanced conversion of amino acids into glucose, which has previously been reported with
529 glutamine (247) and is presumed to fuel enhanced gluconeogenesis. However, it could also be
530 indicative that serine is more robustly incorporated into ceramides in mice or animals with
531 diabetes or the metabolic syndrome. Indeed, ceramides are elevated in individuals with
532 metabolic syndrome or diabetes (discussed below), and kinetic studies using stable isotope
533 palmitate tracers have shown that individuals with insulin resistance have increased rates of
536 While sphingolipids are of course consumed in the diet, most are thought to be degraded
537 in the intestine or by the gut microbiome (249, 250). Sphingoid bases are exceptions, as they
25
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
538 can be absorbed by enterocytes (251). Moreover, recent studies indicate that sphingolipids
539 produced by the microbiota can appear in the circulation and peripheral tissues like the liver
540 (252). Additional work is needed, however, to determine the extent to which dietary
541 consumption of sphingolipids alters the body’s lipidome or its susceptibility to disease.
542 The Gonzalez lab found that ceramides and other sphingolipids are resynthesized in the
543 intestine and trafficked to other tissues in the body. The accomplished group published a series
544 of elegant studies revealing that intestinal farnesoid X receptor (FXR) stimulates biosynthesis of
545 enterocyte ceramides by upregulating genes in the de novo synthesis pathway. The newly-
546 generated ceramides then traffic to the liver, where they alter liver metabolism in ways that
547 contribute to hepatic steatosis and insulin resistance (253). Selective antagonists of intestinal
548 FXR lower ceramides and prevent metabolic complications (discussed in greater in the
549 MAFLD/MASH section below)(7, 254-258). The sphingomyelinase nSMase2 is a FXR target
550 gene that links FXR to increased C16:0 ceramides in the intestine and circulation (255).
551 In addition to FXR, HIF2α, a hypoxia stabilized transcription factor, has been shown to
552 control both intestinal and circulating ceramides in obesity. Disruption of Hif2α in the intestines
553 of mice reduces levels of both intestinal and circulating ceramides. These knockout mice
554 showed decreased body weight, improved insulin sensitivity, and resolved hepatic steatosis
555 (259). Hif2α is also expressed in hypoxic white adipose tissue (WAT) from mice, where it
556 reduces the transcription of the alkaline ceramidase encoded by Acer2 to increase ceramides
557 (260).
558 A third intestinal regulator of ceramides in obesity is the transcription factor Myc, which
559 induces Cers4 transcription to further increase levels of circulating ceramides (261(261).
560 Treating mice with gut-targeted Myc inhibitors increased glucagon-like peptide-1 and reduced
561 serum ceramides, which improved metabolic outcomes. While intestinal FXR, HIF2α, and MYC
26
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
562 all influence metabolic outcomes in different ways, they may share a common mechanism
563 involving the increased transcription of genes that drive ceramide production.
565 As noted above, the earliest reports that ceramide was a signaling entity resulted from
566 analyses of the inflammatory pathways linking TNFα action to the induction of apoptosis (8, 73-
567 80). In the years that followed, numerous other inflammatory agonists were shown to increase
568 cellular ceramide levels, owing to their ability both acutely activate sphingomyelinases and
569 chronically induce expression of de novo synthesis enzymes. Indeed, unbiased lipidomic
570 screens revealed that inflammatory insults selectively upregulated sphingolipids, while having
571 modest effects on most glycerolipids (262). In human serum from insulin resistant subjects,
572 plasma ceramides correlate with circulating cytokine levels (262, 263). Since obesity is
573 associated with increased systemic inflammation (264), the combined increase in saturated fats
574 and inflammatory agents likely leads to synergistic enhancement of ceramide production to
576 Toll-like receptor-4 (TLR4), a pattern recognition receptor (PRR) involved in innate
577 immunity, is a particularly robust inducer of ceramides. The receptor is highly responsive to
578 lipopolysaccharide (LPS), viral proteins, polysaccharides, and other ligands. Curiously,
579 saturated fatty acids were also identified as TLR4 ligands (266), but studies by Febbraio and
580 colleagues suggested that they instead functioned by amplifying receptor signaling (267).
581 Regardless of the precise mechanisms linking saturated fats to TLR4 activation, the receptor is
582 essential for saturated fat-induced stimulation of ceramide synthesis (4, 266), as it regulates
583 transcription of several enzymes in the de novo ceramide biosynthesis pathway (4, 265, 268)
27
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
585 Ceramides may themselves augment inflammation to produce a vicious cycle of
586 ceramide production. For example, ceramides activate the Nod-like receptor (Nlrp3)
587 inflammasome, which promotes cytokine production and secretion (269). One could envision
588 that this positive feedback mechanism would further enhance cytokine-stimulated ceramide
591 Receptors within the progestin and adipoQ (PAQR) superfamily, including the
592 adiponectin receptors ADIPOR1 and ADIPOR2 (also known as PAQR1 and PAQR2), possess a
593 ligand-gated ceramidase activity. Using a yeast system, Villa and colleagues (270) demonstrate
594 that PAQRs liberated the free sphingoid base and that ceramidase inhibitors nullified PAQR
595 signaling (271). Subsequent studies in mice confirmed that mammalian adiponectin receptors,
596 which share homology to the other PAQR receptors, were also ceramidases that were activated
597 by ligand binding (2, 190). The discovery that certain PAQR receptors were ceramidases
598 suggested an exciting means by which hormonal agonists could acutely regulate ceramide
599 levels to affect a wide range of biological responses. In particular, it suggested mechanisms
600 linking the adipokine adiponectin, which signals through PAQR-like receptors, to its broad
601 spectrum of an anti-inflammatory, anti-atherogenic, and anti-diabetic activities (272). This idea
602 challenged the long-prevailing idea that adiponectin signaled through the serine/threonine
604 deprivation (273). Subsequent studies confirmed that AMPK was dispensable for adiponectin
606 Scherer and colleagues were the first to demonstrate that adiponectin, via its cognate
607 receptors, deacylated ceramide and induced formation of the sphingoid base (2). Critically,
608 mutation of histidine residues within the predicted metal-dependent hydrolase domains of
609 AdipoR1 or AdipoR2 prevented ceramidase activity (2). Crystal structures later confirmed that
28
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
610 these zinc-binding residues are contained within a hydrophobic binding pocket that is
611 homologous to known ceramidases and are crucial for adiponectin signaling (274, 275).
612 Subsequent studies with purified receptors, which included a determination of its crystal
613 structure in the presence of exogenous ceramide, confirmed that the receptors were bona fide
614 ceramidases. These studies by Granier and colleagues revealed that adiponectin receptor
615 crystals isolated in the presence of ceramide contained a product of the ceramidase reaction: a
616 bound fatty acid (190). They also demonstrated that the purified receptor had ceramidase
617 activity which was upregulated 25-fold when adiponectin signaled through its receptor. Further
618 supporting the ligand-gated nature of these receptors, mouse models overexpressing AdipoR1
619 or AdipoR2 fail to degrade ceramide when crossed into an adiponectin-deficient background
620 (276). These data confirmed that the receptors were ceramidases and revealed an exciting new
621 paradigm of receptor biology: ligand-activated ceramidases translate information about the
623 Recently, the Pezzolesi and Holland research teams characterized a human adiponectin
624 gene variant. The variant was found in a genetic screen of siblings that were obese, diabetic,
625 and had rapid kidney failure. The frameshift mutation created an adiponectin moiety that lacked
626 the C-terminus of the protein, leaving the collagen-like tail of the protein without the globular
627 head that interacts with the ADIPOR. The mutation was found to be dominant-negative and
628 resulted in highly reduced amounts of high molecular weight adiponectin complexes. Siblings
629 with the mutation had significantly higher serum ceramide concentrations and had worse health
630 outcomes (277). This study demonstrated that—in humans—adiponectin signaling was
632 Other members of the PAQR family also have ceramidase activities. They are part of an
633 extensive superfamily of enzymes with conserved lipid hydrolase domains (278). Work by Nils
634 Halberg and colleagues identified PAQR4 as a golgi-localized ceramidase with anti-apoptotic
29
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
635 effects in breast cancer cells (279). Genetic studies in humans and mice have linked PAQR9 to
636 alterations in lipid metabolism that resides at a ceramide QTL (280, 281). Intriguingly, this
637 ligand-gated signaling system appears conserved down to plants, where osmotin—a peptide
638 hormone important in drought tolerance—signals through an ADIPOR1 homolog PHO36 (282,
639 283). Osmotin cross reacts with AdipoR1 and can have remarkable nutraceutical effects as an
642 To date, relatively few direct ceramide effectors have been identified. Protein kinase C-
643 zeta (PKCζ), protein phosphatase 2A (PP2A) and inhibitor-2 of protein phosphatase 2A
644 (I2PP2A) link ceramides to altered protein phosphorylation (285). Mitochondrial fission factor
645 (MFF) and cyclic AMP response element binding protein-like 1 (CREB3L1) are newly identified
646 targets of ceramide linked to mitochondrial inefficiency and transcription, respectively (286,
647 287). And certain BCL proteins interact with ceramide-rich membrane domains to increase
648 cytochrome c release and initiate apoptosis (288-294). As incredibly hydrophobic compounds,
649 one should not envision them in the same way as classical second messenger signaling
650 molecules. Rather, ceramides change native membrane environments, and likely initiate
653 Alonso and Goni (295) reviewed the highly unique biophysical properties of ceramides,
654 noting that they “exhibit a number of properties not shared by almost any other membrane
655 lipid.” They discussed their extreme hydrophobicity, their ability to enhance the molecular order
656 and enhance the rigidity of bilayers, their capacity to effect lateral phase separation and create
657 sphingolipid-enriched microdomains, their creation of negative curvature, and their capacity to
658 enhance flip-flop of lipids and increase permeability to proteins. Dingjan and Furterman (296)
30
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
659 further elaborated on the unique attributes of the ‘sphingoid motif’ that distinguishes ceramides
660 and other sphingolipids from the more abundant glycerolipids. This motif, which comprises the
661 first five carbons of the sphingoid long-chain base, interacts with other components within the
662 bilayer owing to its ability to undergo hydrogen-bonding, charge-pairing, hydrophobic, and van
663 der Waals interactions. Importantly, this region is capable of donating hydrogen bonds, a feature
664 not apparent in glycerolipids (296). These biophysical properties enable sphingolipid-enriched
665 membrane domains to dramatically alter membrane properties and influence cellular
666 metabolism and survival. Though more work is needed to understand how these biophysical
667 and signaling events occur at the atomic level, they nonetheless suggest how ceramides
668 contribute in such a substantial way to diverse cellular events and ultimately cardiometabolic
669 pathologies.
670 Because of the unique features of their sphingoid motif—as well as the phosphocholine
671 head group in sphingomyelin—ceramides and other sphingolipids form tight associations with
672 each other and with other lipids to increase the melting temperature of lipid bilayer domains.
673 Sphingomyelin interacts tightly with cholesterol via a hydrogen bond generated by the amide
674 group of the sphingomyelin molecule and the 3-hydroxyl group of cholesterol (297).
675 Sphingomyelin also interacts with cholesterol through hydrophobic van der Waal interactions of
676 the cholesterol sterol ring and the ceramide portion of the sphingomyelin molecule. Cholesterol
677 fills gaps between the larger and sterically bulkier sphingomyelins, and as a result these lipid
678 partners separate from phospholipids and form specific domain structures, commonly called
679 lipid rafts (298-300). These highly ordered and tightly packed membrane microdomains confer
680 detergent resistance to membranes, increase the bilayer’s melting temperature, and serve as
682 Ceramides can also interact with each other to form gel-like domains within membranes
683 and/or lipid rafts (301). They can also form tight associations with sphingomyelins and can
31
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
684 displace cholesterol from its natural membrane binding partner to alter membrane fluidity (302-
686 studies with model systems suggest that ceramides induce coalescence of pre-formed lipid rafts
687 to create new signaling platforms (301). Indeed, studies in the late 1990s demonstrated that
688 ceramides—even at low concentrations—make microdomains that can then associate with
689 larger established rafts, ultimately forming platforms that can become as large as 5μm (305-
690 307). Kinoshita and Matsumori suggest that the relative abundance of ceramide and
692 microdomains (301). Thus, ceramide rich domains stabilize the raft and enhance its biophysical
696 metabolic program. They upregulate systems that enhance fatty acid utilization and storage,
697 while slowing pathways that facilitate glucose uptake. The aforementioned ceramide effectors
698 PP2A (308, 309) and PKCζ (310, 311) contribute to this reprogramming by detecting membrane
699 ceramides to alter metabolic signaling. Ceramides activate PP2A by interacting directly with its
700 regulatory B subunit (308), while also relieving its inhibition by I2PP2A (312). They also recruit
701 atypical PKCζ to ceramide-rich caveolae, which alters its ability to associate with and
702 phosphorylate downstream effectors (313). These effects alter protein phosphorylation patterns
704 Sphingolipid-enriched raft domains have been shown to serve as platforms that facilitate
705 fatty acid diffusion through bilayers (314, 315). Indeed, caveolae are particularly abundant in
706 adipocytes (316), a cell type that is subject to periods involving passage of large quantities of
707 fatty acids through the bilayer. Moreover, ceramides induce the translocation of fatty acid
708 translocases (e.g., CD36) from intracellular stores to the plasma membrane, leading to
32
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
709 increased passage of fatty acids through the membrane (317, 318). This latter event involves
710 ceramide activation of PKCζ (318), which has been shown to accumulate in ceramide-enriched
711 membrane domains (311, 313, 319, 320). Ceramides also upregulate SREBP1c-driven
712 transcriptional programs to induce genes that enhance lipid storage (257, 317). The
713 mechanisms underlying this ceramide-SREBP1c axis are uncertain, though PKCζ is a likely
714 intermediate given prior studies implicating atypical PKC isoforms in the regulation of SREBP1c-
715 driven transcription (321). Lastly, ceramides—via PP2A—inhibit hormone sensitive lipase to
716 slow lipolysis, such that fatty acids remain in inert triglyceride depots (197, 322).
717 While ceramides are upregulating fatty acid uptake and sequestration, they also block
718 the insulin-dependent uptake of glucose into skeletal muscle and adipose tissue. Ceramides
719 achieve this metabolic adaptation by inhibiting activation of AKT (323, 324), an anabolic and
720 pro-survival serine/threonine kinase that stimulates the translocation of the GLUT4 glucose
721 transporter from intracellular stores to the plasma membrane. Insulin and other growth factors
724 triphosphate (PIP3), which recruits AKT to the plasma membrane by binding its pleckstrin
725 homology (PH) domain. Once bound by PIP3, AKT gets phosphorylated T308 and S473 (for the
726 AKT1 isoform) by phosphoinositide-dependent protein kinase 1 and the Rictor-bound form of
727 mammalian target of rapamycin (mTOR) (325-330). Ceramides block AKT phosphorylation and
728 activation by two independent and fully separable mechanisms (331). Their effector PP2A
729 dephosphorylates these activating residues (331). They also block translocation of AKT via
730 PKCζ (319, 320, 332), which phosphorylates an inactivating T34 residue within the AKT PH
731 domain (313, 319). GLUT4 translocation can be restored in ceramide-treated adipocytes by
732 overexpressing myristylated, constitutively active forms of AKT (323). The actions of ceramides
733 in these pathways have profound consequences in multiple metabolic diseases and these
33
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
734 mechanisms will be discussed more in depth in our discussion of the role of ceramides in
735 diabetes.
736 Ceramides Diminish Mitochondrial ATP Production and Induce Reactive Oxygen Species
737 Ceramides impair mitochondrial function by diverting electrons in the electron transport
738 chain into reactive oxygen species instead of using them for oxidative phosphorylation, thus
739 diminishing the organelle’s ability to generate ATP (333, 334). Long-chain C16 and C18
740 ceramides influence mitochondrial energetics, while very-long chain ceramides do not. For
741 example, treating isolated mitochondria with C16-ceramide inhibits complex IV (335). By
742 contrast, very-long chain ceramides (e.g. C24 or C24:1) fail to elicit mitochondrial defects (335),
743 In primary cultured hepatocytes, the overexpression of Cers6, which increases C16-ceramides,
744 induces a comparable diminution in oxidative phosphorylation, though in those studies the site
745 of the defect was at complex II (118). Pharmacological inhibition or siRNA-mediated knockdown
746 of sphingomyelinase synthase-2, which increases long and very-long chain ceramide levels by
747 blocking their conversion to the more abundant sphingomyelin, also impairs mitochondrial
749 ceramidase lowers mitochondrial ceramides and increases respiratory efficiency (337).
750 Studies in animal models further confirmed that long-chain ceramides potently affect
751 mitochondrial ATP production and bioenergetics. Knockout mice lacking Cers6 display
752 increased mitochondrial oxygen consumption in adipose tissue and the liver (117, 338).
753 Similarly, pharmacological inhibition or genetic ablation of Cers1, the muscle-enriched CERS
754 isoform that produces long-chain C18-ceramides, increases muscle respiratory capacity (339,
755 340). By comparison, mouse models that display a decrease in long-chain ceramides, such as
756 knockout mice lacking Cert1 or Cers2, display reduced ETC complex activity and increased
34
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
758 Mechanistically, at least a portion of these actions result from ceramide-driven induction
759 of mitochondrial fission (341), which produces smaller and less efficient mitochondria. Brüning
760 and colleagues (338) screened for sphingolipid-binding proteins using SILAC-based proteomics.
761 They employed a bifunctional sphingosine analog pacSph to identify factors that interacted with
762 sphingolipids produced by CERS6, finding that MFF is a binding partner of C16-ceramides. This
763 protein links long-chain ceramides to alterations in mitochondrial morphology and oxidative
764 phosphorylation (338). Moreover, they found that deleting Cers6 in mice or HeLa cells
766 Ceramides also have acute effects on mitochondrial energetics that are unlikely to
767 depend on mitochondrial fission. Gudz and colleagues demonstrated in isolated mitochondria
768 that ceramide analogs interacted directly with complex III of the electron transport chain (333).
769 Kota and coworkers identified a similar association between ceramides and a subunit of
770 complex IV subunit in D6P2T Schwannoma cells (342). Other researchers have speculated that
771 ceramides might alter complex stability (343). Additional work is needed to dissect the precise
772 molecular mechanisms underlying these acute ceramide actions on the electron transport chain.
774 Ceramides contribute to both extrinsic and intrinsic apoptosis. The extrinsic pathway is a
775 receptor-mediated cell death program triggered by the interaction of death ligands with their
776 cognate receptors. These receptors cleave and activate procaspase 8, which in turn acts on
777 downstream caspases that evoke the irreversible progression of the apoptotic cascade. These
778 death receptors—including CD95, CD40, DR5 (TRAIL2), FcγRII, PAF, and CD14—almost
779 invariably activate sphingomyelinases to acutely generate ceramides (344-351). The resultant
780 ceramide-rich platforms cluster the receptors, which is essential for bringing caspases in
781 proximity and initiating the first sequence of proteolytic events. The CD95 receptor, or APO-1, is
782 an illustrative example; once activated by its ligand, FAS, the receptor activates acid
35
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
783 sphingomyelinase to produce ceramide domains that sequester CD95 receptors. These
784 clustered receptors recruit the adaptor protein FADD and caspase 8, which initiates the
785 proteolytic caspase cascade. Within seconds of sphingomyelinase activation, apoptosis ensues
787 Cellular stress stimuli including radiation, UV light, chemotherapies, or infections can
788 activate acid sphingomyelinase to cluster death receptors and trigger extrinsic apoptosis.
789 Genetic knockout or pharmacological inhibition of acid sphingomyelinase blocks the induction of
790 apoptosis by these stimuli (81, 353-359). This mechanism has important implications in cancer
791 biology, and many chemotherapy-resistant cancers contain low levels of ceramides. Cancer cell
792 lines often show reduced expression of sphingomyelinases or de novo synthesizing enzymes
793 such as serine palmitoyltransferase, or in some cases increased expression of ASAH1, CERK,
794 and SPHK1 (360). For example, the colon cancer cell line SW620 was derived from a tumor that
795 was resistant to TRAIL-mediated apoptosis. These cells were found to have significantly
796 reduced ceramides after treatment with the TRAIL ligand (361, 362). If exogenous ceramides
797 were added after TRAIL ligand, the cells underwent normal apoptosis (361). Similarly, Cers6
799 TRAIL, as the absence of ceramides prevented caspase-3 from translocating into the nucleus to
801 The influence of ceramides on apoptosis is not limited to the extrinsic pathway; it also
802 facilitates cytochrome c release from mitochondria to initiate the caspase cascade that drives
803 intrinsic apoptosis. Ceramides traffic from the plasma membrane [298, 299] or the ER (363) to
804 the mitochondrial outer membrane. Alternatively, they may be synthesized directly in
807 mitochondrial fractions (364). As ceramides accrue, they increase the leakage of cytochrome c
36
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
808 (365). Indeed, experimental induction of mitochondrial ceramides is sufficient to promote
809 apoptosis in cultured cell lines. For example, expression of a genetically modified,
810 mitochondrially targeted sphingomyelinase triggers cytochrome c release and apoptosis (366).
812 ceramides to mitochondria and triggers cell death (367). Convincingly, inhibiting enzymes in the
813 ceramide biosynthesis pathway or overexpressing acid ceramidase prevents cell death by
815 The long-chain C16-ceramide is the relevant species that increases mitochondrial outer
816 membrane permeability (MOMP) (8). Several proposed mechanisms link these ceramides to
817 MOMP. Ceramides recruit pro-apoptotic BCL2 family members (e.g., BAX) to mitochondrial
818 membranes (294). These proteins work in concert with ceramides to increase MOMP.
820 protein that enhances MOMP (369). Lastly, Columbini and colleagues have presented
821 compelling data suggesting that C16-ceramide-containing membrane platforms are themselves
822 channels that facilitate transport of cytochrome c (370-373). The channels formation is very
824 incapable of forming these structures and permitting passage of cytochrome c (373, 374).
825 BCLXL1, an anti-apoptotic BCL2 family member, has been shown to perturb these ceramide
826 channels (375-377). As previously described, ceramides also inhibit AKT, which is a pro-survival
829 While ceramides are best known for their role in apoptosis in many cell types, they are
830 key drivers of intestinal stem cell proliferation and survival. Genetic ablation of Sptlc2 (379-381)
831 or pharmacological inhibition of SPT (382) markedly impairs gut health, with the former leading
832 to disruption of the gut architecture and animal death and the latter evoking a profound gastric
37
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
833 enteropathy. Moreover, individuals with inflammatory bowel disease display markedly
834 decreased expression of SPT (379). Mechanistic studies revealed that this resulted from effects
835 within intestinal stem cells, where ceramides alter a metabolic program to enhance proliferation
836 (381). Mice lacking Sptlc2 in stem cells had shorter small intestines, increased inflammation,
837 and permeability, owing to a marked disorganization of the intestinal villi and a nearly complete
838 depletion of the intestinal progenitors form the crypt. Similar findings were observed in
839 Drosophila melanogaster, as overexpression of SPT (lace) and CERS (schlank) accelerated the
840 proliferation of intestinal progenitors, while their knockout markedly slowed rates of cell division.
841 Studies in intestinal organoid 3-dimensional cultures confirmed that ceramides drive
842 stem cell proliferation and survival, as treatment with myriocin or acute depletion of Sptlc2
843 slowed cell proliferation and organoid viability. Organoid crypts could be rescued from cell death
844 with the addition of C2 ceramides. Interestingly, this treatment also reduced necroptotic gene
845 expression suggesting that ceramides may inhibit necroptosis to enhance cell survival in this
846 cell type. Mechanistically, ceramides enhanced stemness by increasing transcription of fatty
847 acid-binding protein 1 (Fabp1) within crypts and not villi (381). FABP1 binds free fatty acids and
848 increases lipid uptake into cells. The increased fatty acid concentrations in ISCs lead to
849 elevated fatty acid oxidation, which has been shown previously to enhance ISC stemness (383-
850 385). Ceramides also increased expression of carnitine palmitoyltransferase CPT1a, a protein
851 that facilitates fatty acid uptake into mitochondria, in crypts of intestinal organoids. Inhibiting
852 CPT1 negated the ability of ceramides to restore viability in Sptlc2 knockout organoids. The
853 ceramide-induced increase of both FABP1 and CPT1a highlights how ceramides are required to
854 maintain ISC vitality. Further investigation demonstrated that Fabp1 expression was linked to
856 that links macronutrient uptake to ISC viability in the intestine. This mechanism could play a role
38
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
857 not only in the maintenance of gut health but also to the increased tumor risk (e.g. human
860 We previously proposed a theory to explain this spectrum of ceramide actions (317). We
861 speculate that ceramides serve as evolutionarily conserved signals of lipid overload, helping
862 cells adjust to the excessive burden of detergent-like FFAs. First, their unique biophysical
864 membranes alter cellular metabolism by decreasing a cell’s reliance on glucose and enhancing
865 fat utilization and storage. As part of this metabolic reprogramming, ceramides decrease
866 mitochondrial efficiency, which allows the cell to decrease the impact of excessive fat
867 metabolism on the electrochemical gradient across the inner mitochondrial membrane. Third, as
868 ceramide levels increase further, they induce apoptosis, which protects the organism from a cell
869 whose membranes have been compromised by the excessive fat load. We presume that stress
870 stimuli such as inflammatory cytokines increase ceramide synthesis as a means of preparing
871 tissues for the increased fatty acids that accompany their stimulation of lipolysis.
872 This hypothetical role for ceramides would explain many of the features of cellular
873 dysfunction that underlies cardiometabolic pathologies, including the decreased glucose
874 utilization, increased triglyceride storage, increased ROS production, and enhanced cell death
875 that characterize the metabolic syndrome, diabetes, and its myriad cardiometabolic
876 complications (Figure 4). Advances in lipidomics further support such roles, as they have
877 revealed strong relationships between serum and tissue ceramides and disease markers in
878 human patients. Indeed, several groups report that circulating ceramides surpass other lipid
879 biomarkers such as LDL cholesterol in diagnosing cardiometabolic disease risk. Below we will
880 explore the roles of ceramides as both causal drivers and biomarkers of cardiometabolic
881 disease.
39
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 4: Progressive Increases in Ceramides Explain the Metabolic
Reprogramming and Subsequent Cell Death that Drives Cardiometabolic
Disorders. As ceramides accrue in aging or disease, they initiate a reprogram
metabolism by inhibiting glucose utilization, increase triglyceride deposition, and
inducing oxidative stress. These are characteristic features of the metabolic
syndrome. As ceramides continue to accrue, they drive the cell death and fibrosis
that defines irreversible features of cardiometabolic diseases.
885 pancreatic islet. This disease is diagnosed clinically by hyperglycemia, driven by the inability of
886 insulin to stimulate the postprandial disposal of glucose in skeletal muscle and adipose tissue,
887 as well as a failure to adequately repress the release of glucose from the liver. Defects in lipid
888 and amino acid homeostasis are also common features of the disease (386, 387).
889 Plasma ceramides have emerged as potential biomarkers of diabetes risk. Moreover,
890 their overaccumulation in liver, muscle, adipose tissue, and pancreatic β-cells contributes to the
40
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
891 pathology by instigating insulin resistance and opposing insulin secretion. As the combined loss
892 of insulin action and insulin production are hallmarks of diabetes, strategies for blunting
894 Emerging Data Suggest that Serum Ceramides may be Biomarkers of Diabetes Risk
895 Multiple clinical studies have shown that circulating ceramides positively correlate with
896 hyperglycemia or other measures of impaired glucose metabolism. The Australian Diabetes,
897 Obesity, and Lifestyle Study, which profiled serum lipids in 640 normoglycemic and prediabetic
898 individuals, revealed that ceramides positively correlated with changes in fasting blood glucose
899 (388). The Strong Heart Family Study, a longitudinal investigation of 2,768 Native Americans,
900 showed that most ceramide species were elevated in aged subjects and associated with higher
901 baseline glucose levels (389). Multiple other studies revealed associations between elevated
902 serum ceramides and a decline in insulin sensitivity, such as homeostatic model of insulin
903 resistance (HOMA-IR) scores. For example, the aforementioned Strong Heart Family Study
904 found that increased serum ceramides, including those comprising 16:0, 18:0, 20:0, and 22:0
905 acyl-chains, correlated with higher plasma insulin and elevated HOMA-IR at baseline and at a
906 subsequent analysis 5.4 years later (390). Interestingly, circulating sphingomyelins—unlike
907 ceramides—were inversely related with plasma insulin, HOMA-IR, and HOMA of β-cell function
908 (390). A study of 2,302 ethnically Chinese Singaporeans similarly found a positive correlation
909 between ceramides and HOMA-IR (391). Interventional dietary studies have also shown that
910 consumption of saturated fat diets negatively impact HOMA-IR while increasing circulating
911 levels of long-chain ceramides (240). When they are measured, dihydroceramides, the
912 precursor of ceramides in the de novo synthesis pathway, are also elevated in the circulation of
913 people with impaired glucose metabolism (240, 392, 393). Thorens and colleagues found that
914 circulating dihydroceramides predicted diabetes susceptibility up to 9 years before the diabetes
915 onset (392). Lastly, in a study of individuals undergoing bariatric surgery, elevated serum
41
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
916 ceramides and dihydroceramides measured at the time of the procedure predicted whether a
919 tissues that control blood glucose levels. An example is the liver, which shows increased rates
920 of gluconeogenesis in patients with type 2 diabetes. In a study analyzing 125 liver biopsies from
921 patients of Finnish descent, hepatic ceramides containing C16:0, C18:0, C22:0, and C24:1 acyl-
922 chains correlated with insulin resistance, independent of body weight or liver triglycerides (240).
923 In a study of adipose tissue, where insulin stimulates glucose disposal and represses lipolysis,
924 multiple liver ceramides correlated with insulin resistance independent of obesity (394). In
925 skeletal muscle, the major site of postprandial glucose disposal, C18:0 ceramides were
926 elevated in insulin-resistant (395) or type 2 diabetic (396) patients independent of obesity.
927 Moreover, Goodpaster and colleagues reported that muscle ceramides associated with insulin
928 resistance in many different study cohorts (397-401). These correlational studies in humans
929 underscore the potential utility of ceramides as markers of diabetes risk, as well as their
931 Ceramides Antagonize Insulin Signaling and Insulin Secretion to Disrupt Glucose
932 Homeostasis
933 Once secreted from the pancreatic islet, insulin binds to insulin receptors (IR) that are
934 abundant in a variety of tissues including muscle, liver, and adipose. IRs are tyrosine kinases
935 that phosphorylate insulin receptor substrates (IRS), which in turn recruit PI3K to initiate protein
936 phosphorylation cascades that promote the uptake and storage of glucose and other nutrients.
937 AKT is a critical intermediate in this signaling cascade that is a central repressor of
938 gluconeogenesis (in the liver) and activator of insulin-stimulated glucose uptake (in muscle and
939 adipose tissue) (further reviewed in (402)). It is also a pro-survival enzyme that prevents
940 apoptosis in a variety of cell types, including the pancreatic β-cells (402). By blocking AKT
42
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
941 activation (discussed above), ceramides play a causal role in T2D by impacting its stimulation of
942 glucose uptake and its anti-apoptotic effects in the pancreatic beta cell (Figure 5).
943
945 Roger Unger’s laboratory, which was studying the mechanisms linking exogenous palmitate to
946 apoptosis in pancreatic islets cultured from Zucker diabetic fatty (ZDF) rats. Ceramides
947 increased in palmitate-treated islets, and exogenous ceramide analogues killed this fragile cell
948 population (403). Fumonisin B1, an inhibitor of the CERS enzymes (Figure 1), prevented
949 palmitate-induced apoptosis of islets, indicating that a sphingolipid was an obligate intermediate
950 in this lipoapoptotic pathway. A follow-up study by Unger’s group revealed that long-chain fatty
951 acids induced expression of SPT genes in the islets of ZDF rats (404). Moreover, they found
43
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
952 that L-cycloserine, a broad spectrum antibiotic that inhibits SPT to blunt de novo ceramide
954 The same year that Unger was publishing these exciting findings, we reported that
956 activation of AKT in 3T3-L1 adipocytes (323). Thus, these parallel studies in the late 1990s
957 revealed potential roles for ceramides as both inducers of insulin resistance and as drivers of β-
958 cell death. A few years later, Poitout and colleagues reported that ceramides also repress
959 insulin gene transcription in cultured insulinoma cells (405), suggesting an additional means by
961 Conclusive evidence that ceramides were essential drivers of insulin resistance and
962 hyperglycemia was provided by our group, when we employed pharmacological and genetic
963 approaches to selectively block ceramide synthesis in diet-induced obese mice as well as
964 Zucker fa/fa and ZDF rats. We also conducted studies using a lipid-infusion model to mimic
965 lipid-driven insulin resistance observed in prediabetes, as well as glucocorticoids to model the
966 insulin resistance found in disorders such as Cushing’s disease. Our studies made heavy use of
967 myriocin, which has a high affinity for SPT and irreversibly inhibits the enzyme complex.
968 Myriocin improved insulin sensitivity in all models and protected ZDF rats from fasting
969 hyperglycemia, insulin resistance, and declining β-cell mass (13). Subsequent studies with
970 myriocin confirmed that it prevented or reversed insulin resistance in diet-induced obesity, lipid-
971 infused rats, leptin-deficient mice and rats, and fructose-fed hamsters (4, 13, 15, 322, 406-408).
972 Studies involving the genetic manipulation of de novo synthesizing enzymes confirmed
973 that ceramides were important regulators of insulin-stimulated glucose disposal. For example,
974 genetic ablation of Sptlc2 or Degs1 or overexpression of Asah1 in mice decreases ceramide
975 concentrations and improves insulin sensitivity. For example, genetic ablation of Sptlc2 in white
976 or brown adipocytes or overexpression of Asah1 in either adipose tissue or liver improved
44
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
977 systemic insulin sensitivity and normalized glucose tolerance in diet-induced obese mice (197,
979 The synthetic retinoid fenretinide, which inhibits DES1 in the de novo synthesis pathway,
980 also improves insulin sensitivity and improves glucose disposal (Figure 1). Fenretinide’s insulin-
981 sensitizing effects were initially attributed to its ability to lower circulating retinol-binding protein
982 4 (RBP4), which had been implicated in insulin resistance (409). However, Preitner et al.
983 showed that fenretinide retained efficacy in Rbp4 knockout mice (410), suggesting that it altered
984 insulin sensitivity through an additional mechanism. Curiously, parallel studies revealed that
985 fenretinide was a potent inhibitor of DES1 in cultured cancer cells that lowered levels of
986 ceramides (411, 412). Bikman et al. confirmed that fenretinide depleted ceramides, increased
987 dihydroceramides, and improved insulin sensitivity in mice maintained on an obesogenic diet
988 (14). This finding suggested that fenretinide’s ability to inhibit DES1 and lower ceramides
989 accounted for the drug’s ability to improve insulin resistance. In collaboration with David Kelley
990 and his team at Merck Research Labs, we later confirmed that inducible depletion of Degs1 in
991 the whole body, liver, or adipose tissue restored insulin resistance in leptin-deficient ob/ob mice
992 or mice fed an obesogenic diet (317). These studies reveal a new therapeutic approach for
993 improving glucose homeostasis. Though SPT inhibitors have failed to reach the clinics owing to
994 an on-target gut toxicity (382), DES1 inhibitors could emerge as an alternative therapeutic
995 approach for lowering ceramides to resolve insulin resistance and improve β-cell health. Indeed,
996 inducible depletion of DES1 throughout the body did not produce any of the gut abnormalities
997 observed following pharmacological or genetic disruption of SPT (317). Moreover, fenretinide
998 has been used extensively in the clinics, with few major side effects (413-424).
999 Long-chain C16 and C18-containing ceramides modulate insulin sensitivity, while very
1000 long chain ceramides including C24 and C24:1 ceramides appear to be either protective or
1001 benign. Brüning and colleagues generated whole-body and tissue specific Cers6 knockout mice,
45
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1002 which have diminished levels of C16:0 ceramides in adipose tissue and the liver (117). These
1003 Cers6 knockout mice were refractory to diet-induced obesity, insulin resistance, and glucose
1004 tolerance and showed marked improvements in insulin signaling to AKT and mitochondrial
1005 respiration. In that study, total ceramide levels didn’t change; the depleted C16 ceramides were
1006 replaced by very long chain ceramides comprising C24 and C24:1 acyl-chains. Concurrently,
1007 our group found that mice that are haploinsufficient for the gene encoding CERS2, the enzyme
1008 that produces the C20-24 ceramides, developed worsened hepatic steatosis and insulin
1009 resistance (406). In this model, depletion of Cers2 led to a compensatory upregulation of Cers6
1010 transcripts and increased production of the pathogenic C16:0 ceramides. Thus, ceramide
1011 acylation patterns, rather than total ceramide levels, are the key determinant of insulin
1012 sensitivity. These findings were reinforced in mouse primary hepatocytes, where overexpression
1013 of Cers6—but not Cers2—antagonized insulin signaling and mitochondrial respiration (Figure
1014 6). Building upon these observations, Brüning’s group used the Cers6 knockout mice to
1015 determine that C16-ceramides regulate mitochondrial fission factor (MFF), which contributes to
1016 the insulin resistant phenotype (338). We surmise that ceramide-induced insulin resistance is
1017 multifactorial, resulting from both effects on proximal insulin signaling (e.g., inhibition of AKT)
1019 Other genetic models have highlighted the importance of ceramides and their
1020 metabolites in the context of diabetes. CERS6 also produces C16:0 ceramides, and knockout
1021 mice lacking the enzyme were shown to have reduced body mass, improved insulin sensitivity,
1022 and improved glucose tolerance (425). Similarly, mice lacking Cers1, which produces the C18:0
1023 ceramides that are plentiful in muscle, display improved insulin sensitivity and glucose tolerance
1024 (426). These observations support the emerging conclusion that long-chain ceramides are
1025 pathogenic. Curiously, sphingomyelin synthase-1 knockout mice displayed improved insulin
1026 sensitivity, which was largely attributable to decreased weight gain on a high-fat diet (427). This
46
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1027 outcome was counter to what had been observed in in vitro studies, where knockdown of the
1028 gene led to increased ceramide accumulation resulting in inhibition of insulin signaling and
1029 impaired mitochondrial function (336). Lastly, glucosylceramides have been shown to impart
1030 liver steatosis and insulin resistance, and inhibitors of glucosylceramide synthase (GCS) have
1031 proven efficacious in reversing steatosis, insulin sensitivity, and glycemic control in obese and
1032 diabetic mice (428-431). Knockout mice for the gene that encodes GM3 synthase, which
1033 synthesizes GM3 gangliosides, were also protected from insulin resistance (432, 433). Using in
1034 vitro approaches, we demonstrated that ceramides and glucosylceramides played separable,
1037 These studies point to the importance of ceramides as causal drivers of insulin
1038 resistance and glucose intolerance (Figure 5). Skeletal muscle is the major site of postprandial
1039 glucose utilization, and interventional studies in cultured myotubes, isolated muscle fibers, or
1040 rodents have confirmed that ceramides are obligate intermediates linking saturated fats,
1041 inflammatory cytokines, or glucocorticoids to the inhibition of insulin signaling and the opposition
1042 of insulin-stimulated glucose uptake (13, 434-438). Watt and colleagues demonstrated that
1043 circulating ceramides present in low density lipoproteins contributed to muscle insulin
1044 resistance. They found that LDL particles carrying ceramides were sufficient to induce insulin
1045 resistance, while those that were devoid of ceramides were unable to oppose insulin action
1046 (439). Moreover, ablation of Degs1 from the liver lowered serum ceramides and improved
1047 muscle insulin sensitivity (317). These studies suggest that ceramides trafficked in the
1049 Ceramides oppose insulin-stimulated glucose uptake by inhibiting AKT (Figure 5) (323),
1050 the aforementioned anabolic kinase that stimulates the translocation of the insulin responsive
1051 GLUT4 glucose transporter to the plasma membrane in muscle and adipose tissue (440).
47
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 6: Hepatic ceramides increase fatty acid uptake and triglyceride storage,
enhance gluconeogenesis, and induce mitochondrial dysfunction. Ceramides induce
translocation of CD36 from the ER to the plasma membrane to increase fatty acid import into
the cell. Ceramides also induce Srebp1f expression to increased glycerolipid synthesis. At
the same time, ceramide de novo synthesis gene expression rises, and the increased
intracellular FFA can then be used as substrate to generate more ceramides. As CERS6
genes become more heavily expressed the expression of lipase genes is diminished. This
creates an environment in the cell where the concentration of both glycerolipids and
sphingolipids is rising. Without the means to liberate FFA from the more complex lipid
molecules via lipases, these lipids then must be stored in lipid droplets. Ceramides, as in
other cells, also activate PP2A and PKCζ which inhibit the actions of AKT, effectively
inhibiting gluconeogenesis. They also diminish mitochondrial efficiency, induce mitochondrial
fission, increase ROS, and stimulate the leakage of cytochrome C to initiate apoptosis.
1052 Several groups, including ours, found that ceramide analogs or agents that induce endogenous
1053 ceramide biosynthesis inhibit activation of AKT by a broad range of tyrosine kinase receptors
1054 (320, 323, 378, 441-447). In the majority of these studies (323, 378, 441, 443, 446-448), but not
1055 all (448-450), ceramides blocked AKT without impacting upstream signaling events such as the
1056 enzyme PI3K. We subsequently determined that ceramides inhibited AKT by two separate
1057 mechanisms that could be defined by their relative sensitivity to the PP2A inhibitor okadaic acid
48
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1058 (OA) (331). Via a pathway not sensitive to OA, ceramides block the insulin-stimulated
1059 translocation of the AKT pleckstrin homology (PH) domain to the plasma membrane, an event
1060 that is important for AKT phosphorylation on two activating residues (T308 and S473 in the
1061 AKT1 isoform). Via an OA sensitive mechanism, ceramides trigger the dephosphorylation of
1062 AKT on these two residues by activating PP2A (Figures 3 and 4). The relative importance of
1063 either pathway varies by cell type, though both can be found to operate simultaneously (331).
1064 Ceramides interact with PP2A directly to help form its active heterotrimer structure. This
1065 effect is specific to ceramides, but not other sphingolipids including the dihydroceramides (308).
1066 Ceramides can also act indirectly on PP2A by displacing the regulatory factor PP2A inhibitory
1067 protein I2PP2A (451). Through these actions, they upregulate PP2A activity, leading to
1069 PKCζ is the ceramide-responsive intermediate that inhibits AKT translocation. The
1070 kinase phosphorylates Thr34 in the AKT PH domain, which prevents its binding to
1071 phosphatidylinositol 3,4,5-triphosphate (PIP3, Figure 5) (319). A key feature of this mechanism
1072 is that AKT and PKCζ interact in caveolin-enriched microdomains. Elevated ceramide
1073 concentrations prevents the kinase from moving to the PIP3-enriched domains that foster its
1074 phosphorylation and activation (313). The relative abundance of these caveolin-enriched
1075 structures determines the relative abundance of caveolae in any given cell type.
1076 Ceramides may also block trafficking of GLUT4 to the membrane through additional
1077 mechanisms downstream of AKT. Using cultured myotubes, the Hoehn and James laboratories
1078 determined that mitochondrial ceramides induce insulin resistance by inducing oxidative stress
1079 (452, 453). James and colleagues obtained additional insight into this mechanism using creative
1080 approaches to increase levels of mitochondrial ceramides (337). By applying proteomics, they
1081 found that mitochondrial ceramides altered the subunit composition of electron transport chain
1082 components. This hypothesized role for oxidative stress is consistent with studies from the
49
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1083 Lander laboratory, which demonstrated that antioxidant interventions were insulin sensitizing in
1084 cultured cells and animal models (454). Hoehn and colleagues also demonstrated that the
1085 mitochondrial permeability transition pore (mPTP) was an essential intermediate in L6 myotubes
1086 that linked ceramide analogs to the inhibition of GLUT4 translocation (455). Klip and colleagues
1087 also found that ceramides influenced GTP-loading of RAC, which they deemed a requisite for
1088 GLUT4 translocation (456). Moreover, they found that ceramides influenced sorting of GLUT4
1089 into an insulin responsive compartment containing syntaxin 6 (457). Thus, additional
1090 mechanisms independent of AKT or oxidative stress may contribute to ceramide-induced insulin
1091 resistance.
1092 Beyond these actions in muscle, ceramides also induce gluconeogenesis and steatosis
1093 in the liver (317). Ablation of hepatic Degs1 improves hepatic insulin sensitivity while restoring
1094 insulin-signaling to AKT (317) (Figure 6), likely through the same mechanisms that were
1095 described above. Degs1 ablation also resolved hepatic steatosis, an effect which is partially
1097 SREBP1 is a master regulator of lipogenesis in the liver activated by ceramides (257) and
1098 ceramide synthases (458). Ceramides also induce the translocation of CD36 to the plasma
1099 membrane from the ER allowing for an influx of FFA from extracellular stores (317). These
1100 actions link ceramides to the hepatic steatosis and hepatic glucose overproduction that are
1102 Ceramides Impair Glucose-Stimulated Insulin Secretion and Decrease Insulin Expression
1104 The islets of Langerhans of the exocrine pancreas comprise multiple cell types, including
1105 α-cells that produce glucagon, β-cells that secrete insulin, and δ-cells that produce
1106 somatastatin. High concentrations of glucose and free fatty acids induce apoptosis in β-cells,
50
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 7: Ceramides in Beta cells. The increase in beta cell ceramides can lead to
increased apoptosis. Ceramides also influence insulin production and secretion by inhibiting
gene transcription, increasing ER stress, impariting electron transport chain (ETC) complex
activity, and diminishing processing of proinsulin in the Golgi.
1107 leading to hypoinsulinemia and frank hyperglycemia (Figure 7)(460). When exposed to excess
1108 nutrients, the β-cell initially attempts to store the excess carbons provided by these fuels as
1109 triglycerides, which form inert lipid depots that protect cells from cytotoxic, detergent-like fatty
1110 acids. However, as the triglyceride depots fill, the excess fatty acids become substrates for the
1111 SPT and CERS enzymes, leading to the biosynthesis of ceramides and other sphingolipids
1112 (461-463). Inflammatory cytokines such as TNFα, IL-1β, IFN-γ, and amyloid and islet amyloid
1113 polypeptide further induce ceramides (464). As discussed above, the Unger lab demonstrated
1114 that ceramides increased islet apoptosis, and that inhibition of these ceramide-synthesizing
1115 enzymes preserved β-cell viability in islets exposed to palmitate (403, 404). Since that initial
51
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1116 discovery, several of the glucolipotoxic and inflammatory agents listed above have been shown
1118 In cultured INS-1 beta cells, glucose and fatty acids not only fuel the sphingolipid
1119 pathway with substrate, but they also increase expression of transcripts encoding Cers4.
1120 Treatment with fumonisin-B1 or siRNA targeting Cers4 transcripts reduced the apoptotic effect
1121 (465). Manukyan et al. reported similar findings in MIN6 cells, finding that myriocin, fumonisin-
1122 B1, or siRNAs that targeted ceramide-synthesizing enzymes prevented palmitate-induced cell
1124 Studies in cultured rat islets have also shown that glucose-stimulated insulin release can
1125 be modulated by ceramides. Kelpe et al. demonstrated that both palmitate and ceramides
1126 reduced preproinsulin mRNA transcription (Figure 7), thus diminishing the amount of available
1127 hormone to respond to postprandial glucose surges. In the palmitate-treated cells, inhibitors of
1128 de novo ceramide synthesis restored insulin transcript levels (405). By contrast, the addition of
1131 Ceramides may also elicit other actions that impact β-cell health and reduce insulin
1132 secretion. Biden and colleagues demonstrated that ceramides are also essential for palmitate-
1133 induced ER stress (468). Ceramides impair mitochondrial respiration, an essential component of
1134 glucose-stimulated insulin secretion (Figure 7) (469, 470). Lastly, Kowluru and colleagues also
1135 found that the small GTP-binding protein Ras-related C3 botulinum toxin substrate 1 (RAC1)
1136 (470-472) and the atypical PP2A subunit ALPHA4 (473) were ceramide targets that contributed
1138 Palmitate is not the only FFA to affect ceramide action. Manukyan et al. demonstrated
1139 that oleate attenuates the effect of palmitate in β-cells of cultured human islets and cultured
52
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1140 MIN6 cells (466). Castell et al. further demonstrated that oleate induced beta cell proliferation in
1141 rats (474). Treatment with myriocin or sphingosine kinase inhibitors decreased oleate-induced
1142 cell proliferation, suggesting that the oleate-derived free fatty acids need to be esterified into a
1143 sphingolipid to elicit these actions. They also knocked down acyl-CoA-binding protein, which
1144 binds medium and long chain esters for transport, and the fatty acid elongase ELOVL1.
1145 Interestingly, the knockdown of either enzyme inhibited oleate’s induction of cell proliferation
1146 (474). This suggests that very long chain sphingolipids may play a protective role and improve
1148 Recent work by Griess et al. further suggest a role for very long chain ceramides
1149 produced by CerS2 in facilitating insulin maturation. Cers2 ablation from the beta cell blunted
1150 the cleavage of c-protein from pro-insulin (475). Their data suggest that very long chain
1151 ceramides may support the cis-Golgi network during insulin processing. Alternatively, since
1152 ablation typically leads to compensatory induction of CERS6 in other cell types (120, 476, 477),
1153 this intervention could also be explained by induction of C16-containing ceramides or, as the
1157 While most of this review focuses on ceramides as a product of obesity that fuel weight-
1158 driven comorbidities, we also surmise that the accumulation of ceramides may disrupt energy
1159 expenditure or regulate food consumption to exacerbate weight gain or hinder weight loss, and
1160 thus contribute to obesity itself. Indeed, once an individual has become obese, the chances of
1161 losing weight and maintaining a healthy body mass index (BMI) becomes much more difficult
1162 (478). Ceramides may exacerbate this problem by influencing neuronal control of energy intake
1163 and expenditure and/or by directly affecting lipolysis and thermogenesis within the adipocyte.
1164 However, our understanding of the role of ceramides in the hypothalamus or other areas of the
53
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1165 brain is incomplete and is an important emerging area of sphingolipid research. Nonetheless,
1166 we will highlight the emerging data suggesting potential roles for ceramides in the control of
1168 Hypothalamic Ceramides May Influence Insulin Sensitivity, Food Intake, and/or Energy
1169 Expenditure
1170 The hypothalamus is a structure within the forebrain that coordinates homeostatic body
1171 processes such as hunger, temperature, energy balance, and sleep. It comprises many different
1172 neuronal populations organized into distinct nuclei (479), some of which have the capacity to
1173 sense hormones and nutrients and translate changes in insulin, glucose, lipids, and amino acids
1174 into alterations in food intake, insulin sensitivity, and energy expenditure (480, 481). In rodents
1175 fed high fat diets or infused with saturated fat emulsions, ceramides accumulate in the
1176 hypothalamus (4, 482-484), where they may contribute to hypothalamic insulin resistance.
1177 Using cultured neurons, Le Stunff and colleagues found that ceramides linked exogenous
1178 palmitate to defects in insulin signaling, employing a PKCζ-dependent mechanism (408) that is
1179 comparable to the one described above for skeletal muscle. Myriocin or siRNA targeting Sptlc2
1181 Le Stunff and colleagues went on to test the consequences of these ceramide actions in
1182 vivo, finding that intracerebroventricular (ICV) injection of C2-ceramides into obese Zucker rats
1183 decreased hypothalamic insulin sensitivity and impaired systemic glucose tolerance (408).
1184 Lopez and colleagues similarly found that ICV infusion of C6 ceramides into rats increased
1185 C16:0 ceramides in the mediobasal hypothalamus, which correlated with increased
1186 inflammation and elevated ER stress and subsequent reductions in sympathetic nerve activity,
1187 systemic insulin sensitivity, and thermogenic capacity (484). By contrast, ICV infusion of
1188 myriocin improved these metrics and preserved glucose stimulated insulin secretion and β-cell
1189 mass (484). Lastly, Lopez’ group determined that the overexpression of Sptlc1 and 2 in the
54
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1190 ventromedial hypothalamic nucleus (VMH) in the hypothalamus induced ER stress (485), while
1191 downregulation of Sptlc1 resolved ER stress and improved metabolic health (486). Interestingly,
1192 treatment with central triiodothyronine (T3), which is known to signal in the VMH to regulate
1193 thermogenesis and browning of WAT, decreased ceramides in the VMH and resolved ER stress
1194 (485).
1195 The appetite-modulating hormones leptin and ghrelin may also modify hypothalamic
1196 ceramides to help confer their effects on food intake and metabolic signaling (480, 487). For
1197 example, leptin decreases hypothalamic ceramides to induce an anorectic response (480),
1198 while ghrelin induces ceramides to mediate its orexigenic effects (487). In the case of ghrelin,
1199 ICV-injection of myriocin negated its ability to stimulate appetite and normalized levels of the
1201 Glycosylated ceramides may also regulate leptin signaling, as over-expression of Ugcg—the
1202 gene encoding GCS—in the arcuate nucleus of the hypothalamus induced weight loss.
1203 Moreover, Ugcg depleted neurons displayed defective leptin receptor signaling (488). Lastly,
1204 targeting Ugcg in the neurons of the mediobasal hypothalamus led to changes in the regulation
1205 of norepinephrine in adipose depots leading to changes in fasting-induced lipolysis (489). These
1206 data suggested that hypothalamic ceramides and glucosylceramides may modulate both
1208 Ceramides may also have roles in non-neuronal cell types within the brain. Increasing
1209 ceramides in astrocytes leads to increased expression of AgRP, the aforementioned leptin-
1210 inhibitable neuropeptide that increases hunger (490). Tanycytes, which line the third ventricle of
1211 the brain and maintain close contact with the hypothalamus, have also been implicated in
1212 trafficking ceramides to multiple nuclei throughout this brain region (491).
1213 As Ceramides Continue to Accumulate in the Brain, they May Contribute to Obesity-
55
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1215 Obesity and T2D increase the risk for neurodegenerative disorders (492, 493), and the
1216 accumulation of ceramides in the brain may also explain this relationship. Alzheimer’s patients
1217 and mouse models display an increase in ceramide content throughout the brain (494, 495).
1218 Ceramides are elevated in amyloid beta plaques in human cadaver samples and rodent models
1219 of the disease (496). Serum ceramides have also been reported to be a biomarker of
1221 ceramides and help resolve behavioral phenotypes associated with Alzheimer’s Disease (498).
1223 Adipocytes, in addition to having an enormous capacity to store energy, sense and
1224 communicate the status of their fuel reserves to the rest of the organism (499, 500). For
1225 example, in relation to their size and/or metabolic activity, adipocytes secrete lipids, peptides,
1226 and adipokines that alter functions of the central nervous system, liver, muscle, heart, etc. The
1227 existence of various adipocyte types (i.e. white, brown, and beige) that differ in the nature of the
1228 lipid droplet, mitochondrial capacity, and thermogenic potential underscores the complexity of
1229 the tissue. Impairments in this energy sensing and storing system are a critical feature of
1230 metabolic disease pathology (501). The accumulation of ceramides in this vital organ can
1231 disrupt these cellular processes to worsen obesity and impair metabolic health (322, 394, 502-
1232 504).
1233 Treating mice and rats with myriocin decreases white adipocyte size, increases numbers
1234 of beige adipocytes, and recruits M2 macrophage into subcutaneous adipose depots (322). To
1235 determine whether this resulted from cell autonomous effects, several groups developed animal
1237 • Scherer and colleagues overexpressed ASAH1, the acid ceramidase, which reduced
1238 levels of C16:0 and C18:0-containing ceramides in both subcutaneous and visceral WAT
56
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1239 depots (276). The intervention also lowered C16:0 and C18:0 ceramide levels in the
1240 liver, leading the authors to speculate that ceramides traffic between the two tissues.
1241 Adipocyte-specific ASAH1 overexpression improved systemic, adipose, and liver insulin
1243 • We depleted Sptlc2 from mouse adipocytes. Knockout of this critical SPT subunit
1244 decreased levels of multiple sphingolipid species, reduced adipocyte size, improved
1245 insulin sensitivity and glucose tolerance, increased energy expenditure and adipose
1247 • Lastly, we selectively depleted Degs1 from mouse adipocytes, finding that it lowered
1248 levels of ceramides and several other sphingolipids, but replaced them with
1249 dihydrosphingolipids containing the canonical double bond in the sphingoid base. Once
1250 again, lowering adipose ceramides decreased adipocyte size, increased adipose tissue
1251 glucose uptake, and enhanced mitochondrial respiration (317). Depletion of Degs1 also
1252 improved glucose tolerance and insulin sensitivity while reducing hepatic steatosis. In
1253 these studies, exogenous ceramides were shown to block lipolysis in adipocytes, thus
1254 restricting FFAs from leaving their preferred storage depot (317).
1255 Supportive data revealing a role for ceramides in adipose health come from whole-body Ormdl3
1256 knockout mice, which display elevated levels of ceramides in adipose tissue, insulin resistance,
1257 and weight gain. Knockout of this SPT suppressor also inhibited brown adipose tissue
1258 thermogenesis and white adipose beiging following exposure to cold or β3 adrenergic agonists
1259 (505). Additional support comes from genome-wide association studies, as ORMDL3 emerged
1261 We note some interesting differences in these knockout mouse studies. While all the
1262 ceramide-lowering interventions mentioned above improved glucose tolerance and resolved
57
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1263 hepatic steatosis, only the myriocin, ORMDL3, and SPTLC2 interventions induced uncoupling
1264 protein-1 (UCP) to effect white adipose browning. By comparison, neither ASAH1 nor DES1
1265 depletion altered adipose thermogenesis. The reasons are unclear, but it suggests that
1266 intermediates in early steps in the pathway, or perhaps the CERS enzymes themselves,
1267 account for these thermogenic actions. Indeed, the CERS enzymes have been identified as
1268 potential transcriptional repressors (127, 507) that are influenced by flux through the pathway.
1269 We also note that two groups that two groups observed a radically different phenotype
1270 resulting from depletion of either the SPTLC1 or SPTLC2 subunits (508, 509). Both groups
1271 observed lipodystrophy owing to impairments in the adipose depot. We previously speculated
1272 that this discrepancy was due to differences in the promoter used to achieve gene depletion,
1273 and that earlier depletion led to impairments in adipose differentiation (322). However, additional
1275 A relatively small number of clinical studies have reported alterations in the
1276 sphingolipidome on adipocytes obtained from obese individuals. In general, human adipose
1277 depots are enriched in ceramides containing the C16:0 acyl-chain (510). Curiously, WAT also
1278 contains disproportionately high levels of ceramides built on non-classical sphingoid backbones
1279 (510), such as deoxyceramides (i.e., ceramides containing a sphingoid base derived from
1280 alanine or glycine versus the more common serine)(510). Profiling studies have revealed that
1281 adipose ceramides are sometimes altered in both obese patients or patients with metabolic
1282 dysfunction. Blachnio-Zabielska et al. found that obese patients had elevated ceramides both
1283 subcutaneous and visceral WAT depots, and that C16:0 ceramides from subcutaneous WAT
1284 correlated with HOMA-IR (502). Our lab found that C16:0 ceramides were elevated in
1285 subcutaneous, but not visceral, white adipose tissue in a small cohort of obese patients that
1286 also have type-2 diabetes compared to obese non-diabetics (322). Another group found
1287 increased C24:1 ceramide in subcutaneous WAT from obese non-diabetic women diagnosed
58
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1288 with hepatic steatosis (394). Another study found that women with severe obesity had elevated
1289 C16:0 and C18:0 ceramide in visceral WAT as compared to subcutaneous WAT (504). Lastly, a
1290 final group found that WAT from obese patients had elevated sphingadienine based
1291 sphingolipids (a sphingoid backbone with two double bonds verses the single double bond in
1292 ceramides) (510). Additional work is needed to identify the relationships between adipose
1294 Much of the work thus far has suggested that C16-containing ceramides are the likely
1295 bioactive species that influence metabolic health. Beyond the profiling data in mice and humans,
1296 Turpin and colleagues found that WAT CERS6 correlates with BMI, adipocyte size, circulating
1297 leptin, systemic insulin resistanc,e and HbA1c levels (117). Moreover, they found that knocking
1298 out Cers6 in whole-body, brown adipose tissue, or liver reduced C16:0 levels in mice, protecting
1299 them from diet-induced weight gain. These mice also had increased lipid utilization in brown
1300 adipocytes and liver and had increased energy expenditure and improved glucose tolerance.
1301 The expression of CerS6 has also been targeted by antisense oligonucleotides in obese leptin
1302 deficient ob/ob mice, which led to a significant reduction in liver and plasma C16:0 ceramides,
1303 lower body weight gain with reduced whole-body fat, and reduced fed and fasted blood glucose
1305 A final study, also focused on brown adipose tissue, further suggests that ceramides
1306 disrupt thermogenesis. BAT is distinct from WAT in that it has a high mitochondrial density with
1307 the ability to dissipate energy and produce heat (197). We found that β-adrenergic agonists
1308 reduce expression of both Sptlc2 and CerS6 in mouse brown adipose tissue, and one of the
1309 ways that β-adrenergic agonists reduce obesity is by lowering BAT ceramides. We confirmed
1310 this mechanism by testing the consequences of BAT-specific Sptlc2 deletion. These mice were
1311 protected from weight gain during high fat diet feeding because of increased systemic energy
1312 expenditure (197). We also tested the effects of BAT specific knockout of Asah1, which
59
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1313 increased ceramides in the tissue by slowing their degradation. These mice had worse
1314 outcomes on high fat diet and showed increased weight, hepatic steatosis, and insulin
1315 resistance. Increasing BAT ceramides disrupted mitochondrial morphology and function. Even
1316 lean, BAT-specific Asah1 knockout mice on a normal mouse chow diet had disrupted
1319 None of the comorbidities of obesity and diabetes are more prevalent or life threatening
1320 than cardiovascular disease (512-515) . Despite FDA-approved drugs that lower blood glucose
1321 levels, mortality among the diabetic population remains unchanged, and an astonishingly 80%
Figure 8: Ceramides influence many major factors that lead to heart disease.
Increased ceramides have been observed in atherosclerosis and ischemic injury
which leads to myocardial infarction. Ceramides are also implicated in increased
apoptosis resulting in HFrEF and likely alter fuel choice to drive cardiac hypertorphy.
Ceramides also increase the chance of developing multiple disease states that are
implicated in the development of both HFpEF and HFrEF.
60
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1322 of diabetic patients 65 or older die from either heart failure or stroke. Hallmark features of
1323 diabetes and heart failure are altered cardiac metabolism and intramyocardial accumulation of
1324 toxic lipid species (Figure 8) (516-519). Ceramides appear to be amongst the most deleterious,
1325 as they are prognostic biomarkers of disease risk and are likely to play causal roles in the
1326 cardiac and vascular dysfunction that underlies these fatal conditions.
1327 Ceramide-Based Scores are in Clinical Use as Measures of Cardiovascular Disease Risk
1328 Several clinics have started measuring circulating ceramides as a means of identifying
1329 patients at risk of cardiovascular disease, as circulating ceramides have been reported to be
1330 better markers of future cardiovascular events than more conventional markers such as LDL-
1331 cholesterol (10, 520). The Cardiac Event Risk Test 1 (CERT1) developed by Zora Biosciences,
1332 which has been licensed by Quest Diagnostics and is in use at several hospitals in the USA and
1333 Europe, is reported to be an effective measure of 5-year risk of cardiovascular mortality. CERT1
1334 incorporates individual C16:0, C18:0, and C24:1 ceramides, as well as the ratio of these three
1335 species to C24:0 ceramide, to create a 12-point score that denotes future likelihood of having a
1336 major adverse cardiac event (MACE)(521). Zora researchers subsequently developed a CERT2
1337 score that incorporated phosphatidylcholine species to measure sample stability (522). We used
1338 a machine learning approach to assess coronary artery disease and developed a sphingolipid
1339 inclusive score (SIC) that included various non-abundant species, including some ceramides,
1340 that outperformed standard metrics of cardiovascular disease risk and coronary artery disease
1341 and which forecasted disease severity (523). Below we discuss the series of studies that
1343 Advancements in lipidomics and mass spectrometry, which enabled relatively high
1344 throughput analysis of clinical specimens, were critical for the discovery that ceramides are
1345 biomarkers of cardiovascular disease. Laaksonen et al. profiled lipids from the Corogene study,
1346 which analyzed samples from more than 5300 patients receiving coronary angiograms at the
61
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1347 Helsinki University Central Hospital. These investigators reported that three ceramide species
1348 (C16:0, C18:0, and C24:1) positively associated with major adverse cardiovascular events.
1349 These ceramide species were associated with the worst possible outcomes including death,
1350 while C24:0 ceramide showed an inverse relationship (521). This research group went on to
1351 validate this ceramide score in the FINRISK study composed of individuals that were monitored
1352 by blood collection for 14 years. They reported that the addition of ceramide C18:0 to other
1353 classical clinical markers such as HDL and LDL improved the prediction of MACE risk. The
1354 ceramide score predicted MACE in the entire FINRISK population (524). In all, this group
1355 published ~15 studies using ceramide-based scores that predicted or correlated with
1356 cardiovascular disease endpoints across multiple clinical populations (39). For example,
1357 Mantovani et al. used data from 30,000 patients across seven cohort studies and again found
1358 that ceramide species, particularly C16:0, C18:0 and C24:1, consistently associated with
1359 cardiovascular disease (525, 526). More recently, Leiherer tested the prognostic utility of this
1360 score in 379 patients with peripheral artery disease, finding that elevated CERT scores
1361 predicted all-cause mortality at 10-year follow-up (527). To date, Zora has explored these
1362 relationships between ceramides and cardiovascular disease in greater than 100,000 subjects.
1363 The relationships between ceramides and cardiovascular disease have also been
1364 reported by other study groups. Schaffer and colleagues profiled ceramides in the Framingham
1365 Heart Study and the Study of Health in Pomerania cohorts, confirming that C16:0 ceramides
1366 were potent markers of disease risk, while C24:0 ceramides had an inverse relationship (528,
1367 529). They similarly proposed that ratios between C16:0 and C24:0 ceramides were a strong
1368 predictor of disease risk. Lemaitre et al. similarly profiled lipids in the Strong Heart Family Study
1369 of Native Americans, demonstrating that elevated C16:0 ceramide and C16:0 sphingomyelin
1370 were associated with increased risk of heart failure, while very long chain plasma C22:0
62
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1371 ceramide and C20:0, C22-0, and C24-0 sphingomyelins were associated with decreased risk
1372 (530).
1373 Ceramide-based scores also show associations with other metabolic readouts, many of
1374 which are associated with increased risk of cardiovascular disease. Studies in the Strong Heart
1375 Study revealed the aforementioned association of ceramides with HOMA-IR and diabetes risk,
1376 two additional risk factors of cardiovascular disease (531-533). Moreover, Jensen et al. showed
1377 that ceramide C16:0 was associated with increased cardiovascular disease risk after the onset
1378 of diabetes (534). A similar study using samples from the Australian Diabetes, Obesity and
1379 Lifestyle Study showed that ceramides were associated with fasting blood glucose, as well as
1380 other cardiometabolic risk factors (388). Thus, ceramides show relationships with the impaired
1382 Additional studies with human heart biopsies have shown that ceramides may
1383 accumulate in the failing myocardium (37). In severe cases of heart failure with reduced
1384 ejection fraction (HFrEF), patients often undergo surgery to receive left ventricular assistance
1385 device (LVAD) which improves cardiac remodeling, function, and metabolism (535-537). LVAD
1388 HFrEF encompasses ischemic and non-ischemic heart failure and is characterized by
1389 dilated ventricles and systolic dysfunction. In the 1990s, it accounted for three-quarters of all
1390 heart failure diagnoses, but is now being overtaken by heart failure with preserved ejection
1391 fraction (HFpEF), which is characterized by diastolic dysfunction and exercise intolerance and
1392 now accounts for the majority of heart failure diagnoses (538). In fact, HFpEF was estimated to
1393 account for 69% percent of all heart failure (HF) cases in 2022 (539) and shows no signs of
63
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1394 slowing, as diagnoses steadily increase at a rate of 1% annually (540). Cardiac ceramides may
1396 HFrEF is often a result of ischemic injury and chronic 𝛽-adrenergic signaling and is
1397 followed by cardiomyocyte apoptosis, inflammation, and subsequent fibrotic repair (516). These
1398 changes ultimately prevent proper contractility of the left ventricle. By contrast, HFpEF is
1399 preceded by chronic morbidities like obesity and leads to cardiac fibrosis in the absence of
1400 severe apoptosis, causing hypertrophy and left ventricular diastolic dysfunction (516). The
1401 diseases share common risk factors. Diabetes increases one’s risk for both forms of the
1402 disease, and both HFrEF and HFpEF patients display similar decreases in nitric oxide (NO) and
1404 Goldberg and colleagues provided the first demonstration that ceramides damaged
1405 cardiomyocytes in vivo. They employed mice with cardiac-specific overexpression of lipoprotein
1407 cardiomyopathy (22). Administration of myriocin to 𝐿𝑃𝐿 overexpressing mice lowered cardiac
1408 ceramides, improved systolic function and prolonged survival rates (22). Similarly, whole-body
1409 heterozygosity for Sptlc1 lowered ceramides and improved cardiac function in 𝐿𝑃𝐿
1410 overexpressing mice. Mechanistic analysis revealed that ceramides upregulated PDK4, atrial
1411 natriuretic peptide, and brain natriuretic peptide: three markers of HF (22). Schulze and
1412 colleagues arrived at a similar conclusion about the effects of ceramides in the heart using an
1413 experimental mouse model of myocardial infraction (i.e., ligation of the left coronary artery).
1414 They found that prior administration with myriocin reduced ventricular remodeling, cardiac
1415 fibrosis, and inflammation (Figure 9) (37). Mechanistically, ceramides have been shown to
1416 disrupt glucose metabolism, induce oxidative stress, exacerbate fibrosis, and drive
1417 cardiomyocyte apoptosis (37, 535, 541-543), and each of these mechanisms could contribute to
64
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 9: Ceramides in cardiomyocytes. Increased influx of FFA into cardiomyocytes and
β-adrenergic stimulation increase de novo ceramide synthesis. Concurrently, cytokine
signaling increases the catabolism of sphingomyelin into ceramides. The increase of
ceramides leads to a loss of mitochondrial efficiency and eventually apoptosis. Ceramides
are also linked to an increase in fibrosis, hypertrophy, and disrupted autophagy.
1419 Studies employing genetic models that allow for cardiac-specific manipulation of
1420 ceramide-modifying enzymes have been few and far between. Sasset et al. found that
1421 cardiomyocyte-specific ablation of the regulatory protein Nogo-A increased SPT complex
1422 activity and exacerbated heart failure caused by transverse aortic constriction (233). This is
1423 consistent with the idea that ceramides are deleterious in the heart, and that the protective
1424 effects of myriocin result from lowering ceramides within the cardiomyocyte. However, Park and
1425 colleagues found that constitutive, cardiac-specific deletion of Sptlc2 induced cardiac
1426 dysfunction in the absence of other treatments (36). Thus, it remains unclear to what extent the
65
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 10: Elevated ceramides leads to endothelial cell dysfunction. Impairment of the
vascular endothelium is largely due to a loss of nitric oxide (NO) production and availability.
Ceramides can reduce NO by inhibiting the activity of endothelial nitric oxide synthase
(eNOS). In the mitochondria, ceramides increase ROS which decrease eNOS activity.
Ceramides also stabilize PP2A and block the inhibitory action of I2PP2A. This results in
PP2A directly interacting with eNOS and AKT to block their interaction, resulting in
diminished NO and impaired vasodilation.
1427 beneficial effects of myriocin are solely attributable to actions within the cardiomyocyte vs. other
1430 The vascular endothelium coats the interior walls of arteries, veins, and capillaries
1431 (Figure 10). It consists of a thin layer of endothelial cells (ECs) that attach to the basal lamina,
1432 structurally building the intima in blood vessels. The ECs are responsible for maintaining
1433 vascular homeostasis by perceiving hemodynamic changes, like shear stress, and blood-borne
1434 substances, like hormones (544). In healthy conditions, ECs control the balance between
66
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1435 vasorelaxation and vasoconstriction, anti- and pro-inflammation, anti- and pro-oxidative stress
1436 by titrating the production of nitric oxide (NO) endothelial NO synthase (eNOS) (545-547).
1439 complications, such as hypertension and atherosclerosis, as well as in many patients with
1442 vascular homeostasis. For example, individuals with the metabolic syndrome frequently display
1444 bradykinin, and blood flow—due to reduced or impaired NO production (550). EC dysfunction is
1445 tightly linked to the development of cardiovascular problems during hyperlipidemic conditions.
1446 Elevation of circulating fatty acids, largely in the form of triglycerides found in circulating
1447 lipoproteins, not only stimulates the buildup of lipid deposits within the blood vessel lumen, but
1448 also promotes formation of ceramides within vascular endothelial cells. This ectopic lipid
1449 accumulation in the vasculature induces inflammatory responses and apoptosis while impairing
1452 phosphorylation on two critical residues: T495 and S1177. Phosphorylation of the S1177 site
1453 activates the enzyme, whereas phosphorylation of T495 is inhibitory (552). As discussed above,
1454 ceramide activates protein phosphatase 2A (PP2A), leading to the dephosphorylation of S1177
1455 (33, 451, 553). Mechanistically, this occurs through an intermediary inhibitor 2 of PP2A
1456 (I2PP2A), which blocks the association of PP2A with its substrates (554). Ceramide disrupts this
1457 association, leading to an interaction between PP2A and eNOS and the dissociation of eNOS
1458 from the activating kinase complex AKT-Hsp90 (33, 451). Blocking ceramide production with
1459 myriocin or Degs1ablation ameliorates high fat-induced endothelial dysfunction (33, 451). This
67
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1460 mechanism highlights the direct ability of ceramides to cause EC dysfunction in the circulatory
1461 system.
1463 phosphate (NADPH) oxidase also influence eNOS activity. In animal models with hypertension,
1464 diabetes, or atherosclerosis, NADPH oxidase levels within the vascular wall are elevated,
1465 leading to high levels of ROS and EC-dependent vascular dysfunction (555-557). Early in vitro
1466 studies have found that synthetic ceramides elevate ROS production, perhaps due to NADPH
1467 oxidase activation, in human and animal EC (558, 559). Indeed, Zhang and colleagues
1468 demonstrated that the ceramide impairment of EC-dependent vasodilation in bovine small
1469 coronary arteries could be attenuated with an NADPH oxidase inhibitor (559). Another study
1470 similarly found that overexpression of CuZn superoxide dismutase, an enzyme that scavenges
1471 ROS, prevents ceramide-driven impairment of EC-dependent vasodilation (560). These findings
1472 suggest that ceramides my also influence EC reactivity through ROS specific pathways.
1473 Multiple studies have also demonstrated that ceramide can alter mitochondria, producing
1474 ROS by disrupting the mitochondrial electron transport chain or inducing apoptosis by altering
1475 mitochondrial outer membrane permeability (561). These actions, discussed previously, may
1476 also prove relevant to EC function, particularly in response to the inflammatory cytokine tumor
1477 necrosis factor-α (TNF-α). TNF-α increases endogenous ceramide content of human and bovine
1478 ECs by activating sphingomyelinase (562, 563). Inhibiting ceramide biosynthesis with the
1479 ceramide synthase inhibitor fumonisin B1 protects bovine cerebral endothelial cells from TNF-
1480 α/cycloheximide-induced cell death (563). This interrelationship between ceramide and TNF-α
1481 suggests that ceramide production is essential for the development of inflammation-induced EC
1482 dysfunction.
1483 Curiously, S1P is a particularly potent regulatory of endothelial cell function and vascular
1484 tone (56, 227, 234, 235), and a disproportionately high percentage of ceramides are likely
68
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1485 converted to this bioactive metabolite in endothelial cells. Di Lorenzo and colleagues found that
1486 disruption of endothelial Nogo-B could restore sphingolipid biosynthesis and the production of
1487 S1P and prevent development of cardiac hypertrophy during pressure overload (235). Similarly,
1488 EC-specific knockout of the S1P transporter SPNS2 prevented this paracrine signaling axis and
1489 induced hypertension (564). Thus, despite the abundance of data implicating ceramides in the
1493 Atherosclerosis is a leading cause of death worldwide and the pathological basis of
1494 many cardiovascular disorders. Multiple factors contribute to its development, including
1495 hyperlipidemia and inflammation. The progression of atherosclerosis is not fully understood, but
1496 it is initiated by lipoprotein retention in the vascular intima. The accumulated lipoproteins mainly
1497 consist of a variety of low-density lipoproteins (LDL), which are too large to pass through EC
1498 barriers. When lipoproteins aggregate in the intima, ECs become activated to recruit monocytes,
1499 resulting in the induction of macrophage chemotaxis and foam cell formation (reviewed more
1500 extensively in (548)). During this progression, ECs generate a large amount of ROS and limit
1501 NO bioavailability, which can exacerbate inflammatory responses while reducing EC-dependent
1502 vasodilation.
1503 In early stages of coronary artery disease, sphingomyelin is elevated in patient plasma
1504 and becomes an independent risk factor of CVD (565). Inhibition of sphingomyelin synthases
1505 (SMS) and reduced production of sphingomyelin has been explored as a potential therapeutic
1506 target for atherosclerosis (566). In mouse models, adenoviral overexpression of both Sgms1
1507 and Sgms2, which code for their respective SMS protein, increases plasma sphingomyelin
1508 levels and worsens atherosclerosis (567, 568). By contrast, Sgms2 deficiency in mice
69
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1510 SGMS1 elicits several abnormalities including metabolic and inflammatory dysfunction (572,
1511 573).
1512 High ceramide concentrations are also found in atherosclerotic plaques and ceramides
1513 are implicated in lipoprotein aggregation, which is the initial step of atherosclerosis (574, 575).
1514 In addition, circulating ceramide species predict plaque instability (576) and detrimental
1515 coronary artery disease outcomes, including future mortality (521, 577). Several studies have
1516 demonstrated that reducing de novo ceramide synthesis can be a potent therapeutic target to
1518 plasma ceramide concentrations and atherosclerotic lesion size in aorta (16, 34, 578) of
1519 atherosclerosis prone ApoE knockout mice fed high fat and cholesterol diet. Moreover, myriocin
1520 treatment reduces plasma cholesterol levels in the same mouse model (578). This phenomenon
1522 Another approach for lowering ceramides is by targeting the enzyme Des1. It has been
1523 demonstrated that inhibition of Des1 using a chemical inhibitor, fenretinide, or genetic
1524 modification in mouse protects the development of vascular dysfunction, though this may be
1525 attributed to its effect on diet-induced obesity, fatty liver, and insulin resistance (13, 14, 33, 317,
1526 580, 581). Surprisingly, fenretinide has been shown to worsen atherosclerosis in ApoE-deficient
1527 mice, though this is likely to be driven by the drug’s retinoid actions, and independent of effects
1528 on Des1 (582). Nonetheless, additional studies are needed to determine the effects of DES1
1532 which range in severity (583). Metabolic dysfunction Associated Fatty Liver Disease (MAFLD) is
1533 a common, relatively benign, and reversible state characterized by excess lipid accumulation
70
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1534 and insulin resistance in the liver that increases one’s risk for major adverse cardiac events
1535 (584). We discussed the means by which ceramides influence liver metabolism in MAFLD in the
1536 preceding sections. In more rare and untreated conditions, MAFLD can progress to more
1537 advanced and deadly Metabolic Dysfunction Associated Steatohepatitis (MASH). MASH,
1538 characterized by fibrosis and inflammation, can lead to hepatic scarring (cirrhosis) which
1539 increases the risk for hepatocellular carcinoma (HCC)— the most common liver cancer and 3rd
1540 leading cause of cancer related death in the United States (585).
1541 The dominant histological and diagnosable feature of steatotic liver diseases is the
1542 accumulation of liver triglycerides, and the liver has an impressive capacity to safely store
1543 excess fatty acids within lipid droplets. However, as these triglyceride stores become saturated,
1544 sphingolipids such as ceramides start to accrue (586, 587). An emerging body of literature
1545 suggests that ceramides contribute to the fibrosis, inflammation, apoptosis, and mitochondrial
1547
1549 Owing to the limited availability of liver biopsies from healthy and diseased livers for both
1550 biochemical assessments and disease confirmation, the literature evaluating relationships
1551 between sphingolipids and NASH severity is somewhat limited. More extensive analyses have
1552 been conducted using blood samples, which are easier to obtain. Herein we discuss the
1553 spectrum of profiling studies looking at liver and blood ceramides in subjects with histologically
71
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Figure 11: Ceramides contribute to MAFLD/MASH. Ceramides have been implicated in
various features of MAFLD. And MASH Decreasing ceramides through pharmacological
inhibition or genetic ablation of de novo ceramide synthesis genes improve outcomes related
to these Steatotic Liver Diseases.
1555 Several small studies suggest that ceramides are elevated in individuals with a liver
1556 disease diagnosis (588, 589). In obese adults, serum ceramides are elevated in patients with
1557 MAFLD and MASH, as compared to serum from patients with histologically normal livers (590).
1558 Wasilewska and colleagues observed similar increases in multiple serum ceramide species in
1559 31 obese children with MAFLD, as compared to 14 controls (591). Serum ceramides and
1560 hepatic triglycerides comparably respond to lifestyle interventions that resolve the disease
1561 burden. For example, Pomrat and colleagues observed lower serum ceramide profiles and
1562 decreased liver SPTLC2 and CERS1 transcripts in 31 histologically confirmed MASH subjects
1563 that underwent a 1-year weight loss intervention (592). Conversely, overfeeding overweight
1564 subjects saturated-fat rich meals disproportionally increased hepatic steatosis, insulin resistance
72
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1565 and ceramides; by comparison, calorically similar overfeeding of high-carbohydrate or
1566 unsaturated fat-enriched diets had more modest effects (240). In the largest study of its kind,
1567 the Orecek lab recently evaluated serum samples from the European MAFLD Registry,
1568 consisting of 627 histologically documented cases of fatty liver disease ranging in severity from
1569 MAFLD to MASH-F4/cirrhosis (593). Interestingly, ceramides were elevated in association with
1570 steatosis, but did not significantly change during the progression to MASH and cirrhosis. By
1571 contrast, sphingomyelins were part of a select group of lipids that changed during these
1572 transitions. Additional work has pointed to select changes in sphingolipids during the transition
1573 between cirrhosis and hepatocarcinoma, as Wang and colleagues found that serum S1P, C16:0
1574 ceramide, and C24:0 ceramides were elevated in patients with hepatocellular carcinoma as
1576 With the expansion of human genetics studies and interests in personalized medicine,
1577 differential causes of fatty liver will likely emerge. Alonso and colleagues identified two distinct
1578 metabolomic subtypes in 535 subjects with MAFLD or MASH, with enhanced accumulation of
1579 sphingomyelin and phosphatidylcholine in the serum of 49% of subjects (596). A strong
1580 example of a genetic cause of this variance is provided by patients expressing a common
1582 patients to MAFLD. Typically, carriers of this variant (rs738409) remain metabolically healthy,
1583 with less insulin resistance, diabetes, and cardiovascular disease, despite the presence of
1584 excessive liver fat (597, 598). The Hobbs and Cohen labs elegantly demonstrated that this
1585 steatotic effect of PNPLA3 is driven by its function as a lipid droplet protein, and triglyceride
1586 accumulation is not reliant on PNPLA3’s enzymatic activity (599). Luukkonen and colleagues
1587 nicely stratified MAFLD cohorts into insulin resistant subjects with or without this common I148M
1588 variant. Liver ceramides were elevated in patients with insulin resistant MAFLD (587), but were
1589 not elevated in individuals with PNPLA3-mediated MAFLD, as this lipid droplet associated
73
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1590 protein more effectively stores fat as triglycerides. Carriers of the PNPLA3 variant still develop
1591 severe fibrosis (598), suggesting the presence of ceramide-independent drivers of collagen
1592 deposition.
1593 Hepatic Ceramides Contribute to the Metabolic Dysfunction that Characterizes MAFLD
1596 hepatocytes (i.e., hepatic steatosis). This steatosis is relatively benign, as it typically presents
1597 asymptomatically, and triglyceride accumulation does not associate with increased mortality
1598 (600). It is present in a fourth of the global population (601, 602). The condition can progress to
1599 MASH, which is further characterized by inflammation, hepatocellular damage, and fibrosis
1600 (Figure 11). MASH afflicts an estimated ~9% of Americans, including a disproportionate
1601 number of males and Hispanics (603). MASH presents a considerable health threat, as it
1602 increases risk of coronary artery disease, diabetes, hepatocellular carcinoma (HCC) and
1603 cardiovascular-related mortality. Despite considerable attention, the mechanisms driving the
1604 transition from MAFLD to MASH are poorly understood. As described below, ceramides may
1606 MAFLD’s transition to MASH is almost certainly multifactorial, with elements intrinsic and
1607 extrinsic to the liver likely contributing. Epidemiological evidence convincingly demonstrates a
1608 relationship between obesity, specifically visceral adiposity, and MAFLD (604). Bidirectionality
1609 between MAFLD and insulin resistance exists, with epidemiological data suggesting that the
1610 transition from MAFLD to MASH is hastened by insulin resistance (605). Interestingly,
1611 ceramides distinguish MAFLD patients with insulin resistance from those with preserved
1612 glucose metabolism (587). Meaning they could serve as biomarkers of the progression of
1614
74
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1615 In preclinical models, interventional studies invariably demonstrate that ceramides are
1616 necessary for MAFLD development. For example, myriocin has been used to resolve hepatic
1617 steatosis (13, 606, 607). We have also observed reduced hepatic steatosis in mice that allow for
1618 the global or conditional depletion of the Degs1 (13, 317). Deleting Degs1 allowed us to acutely
1619 replace tissue ceramides with dihydroceramides. In turn, this prevented or reversed hepatic
1620 steatosis in leptin deficient or high fat fed mice by blunting lipid uptake (e.g. by blocking
1621 translocation of the fatty acid translocase CD36) and decreasing expression of lipogenic genes
1622 (e.g., Srebf1) (317). We have reported similar findings using mice that allow for inducible, liver-
1624 which induce the local degradation of ceramide (318, 608). Additionally, liver-specific deletion
1625 of Cers6 lowered ceramides, normalized lipid uptake, prevented mitochondrial dysfunction and
1626 resolve diet-induced hepatic steatosis(117, 338). Conversely, in vitro overexpression of Cers6 in
1627 primary hepatocytes was sufficient to increase C16-ceramide and triglyceride concentrations
1628 while impairing mitochondrial respiration (118). Blocking ceramide synthesis or stimulating
1629 ceramide degradation are thus complementary approaches for lowering hepatic ceramides to
1632 Ceramides produced outside the liver may play an added role in increasing the hepatic
1633 ceramide burden in MAFLD. Two key tissues, adipose and intestine, shuttle ceramides through
1634 the circulation to the liver to induce steatosis and insulin resistance. Indeed, portal blood flow
1635 coming from the splanchnic bed enriched with visceral fat and intestinal-derived lipids contains
1636 markedly more ceramide in obese rats with MAFLD (13). Strategies for selectively reducing
1637 ceramide accumulation in adipose can also resolve hepatic steatosis, and simultaneously lower
75
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1639 As previously discussed in the obesity section, FXR signaling in the intestine has an
1640 impact on hepatic ceramide concentrations. Intestinal FXR signaling induces ceramide
1641 synthesis. These ceramides then are trafficked to the liver as inducers of hepatic steatosis (7,
1642 257-259). FXR antagonists given to mice lowered ceramides, resolved hepatic steatosis, and
1643 improve insulin sensitivity (7, 257, 258). Subsequently, Xie and colleagues described the
1645 ceramide salvage pathway (259). Ablation of Hif2α from the intestine substantially reduced
1646 HFD-induced obesity and hepatic steatosis in mice, which were very similar to the phenotypes
1647 seen with intestinal FXR ablation. More recently the group has identified roles for Myc-driven
1648 increases in Cers4 and AMPK-driven increases in sphingomyelinase (in response to nicotine)
1649 as drivers of ceramide production in the intestine, which impact hepatic steatosis (261, 610).
1650 Intestinal ceramide production, occurring through multiple unique processes, thus leads to the
1651 enhanced delivery of ceramides to the liver, establishing the gut as a prominent and regulated
1652 source of serum ceramides that can drastically alter hepatic metabolism (611).
1653
1654 SUMMARY
1655 Excessive delivery of lipids to the liver, skeletal muscle, adipose tissue, pancreatic β-
1656 cells, heart, and vasculature induces defects that underlie a broad spectrum of cardiometabolic
1657 diseases. Among the numerous lipid species that accumulate, sphingolipids such as ceramides
1658 are particularly deleterious, as they alter tissue fuel use, decrease mitochondrial efficiency, and
1659 induce apoptosis. Pharmacological or genetic lowering of ceramides by targeting the enzymes
1660 that control their synthesis or degradation resolves insulin resistance, hepatic steatosis,
1661 atherosclerosis, and heart failure in rodents (Table 1). Moreover, ceramides are emerging as
1662 potent biomarkers of a number of diseases, and clinics are now measuring these molecules as
1663 markers of disease risk. Additional investigation of this pathway holds enormous process for
1664 producing new diagnostic tools and therapies for a broad range of cardiometabolic disorders.
76
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1665 Despite the exciting promise of investigations of this interesting class of molecules,
1667 • What are the genetic drivers of hyperceramidemia in humans? Though we’ve described
1668 a handful of gene variants that influence ceramide accumulation, this area of inquiry is
1669 still in its infancy. Many large biobanks that have been used for genetic analysis do not
1670 have accompanying lipidomic data. The identification of gene variants that drive
1672 biology and for stratifying patients for different treatment regimens.
1673 • Could refined subsets of ceramide-based scores better stratify specific cardiometabolic
1674 disease conditions? And how do we advise patients that receive a high ceramide-bsed
1675 score? Though ceramides are now being measured clinically, the complexity of the
1676 sphingolipidome likely offers considerable potential to use subsets of ceramide scores to
1677 better identify tissue defects and disease susceptibility. For example, c18-containing
1678 ceramides are primary produced in skeletal muscle. They are associated with insulin
1679 resistance, the primary site of glucose disposal. This information suggests that an
1680 understanding of acyl-patterns of blood ceramides could unveil sites of biological lesions
1682 patients with high blood ceramides are sorely lacking. Additional interrogation of their
1684 • What are the key intracellular sensors of rising ceramide levels and can we discern
1686 mapping pathways that lie downstream of ceramides, the molecular sensing machinery
1687 remains elusive. A recent study on ceramide activation of CREB3L1 have suggested
1688 that ceramide-enriched membranes flip the orientation of the membrane spanning
77
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1690 This intriguing and attractive mechanism has not been validated by other groups, nor
1691 have scientists determined whether this is a universal means of sending ceramide-rich
1692 platforms. Moreover, it doesn’t explain the other signaling cascades that are downstream
1693 of ceramides, such as the PP2A and PKCζ dependent mechanisms that govern
1694 downstream protein phosphorylation events. Attention to this important question about
1698 discovery efforts have focused primarily on SPT as a therapeutic target, but small
1699 molecule inhibitors of this target have failed to reach the clinics owing to an on-target gut
1700 toxicity (382). One hopes that other targets in the pathway, such as the CERS or DES
1701 proteins or other regulatory interacting proteins that control ceramide flux, could emerge
1702 as safe and useful drug targets. The broad spectrum of disease indications that are
1705 Dedicated efforts to address these questions could have a marked impact on our understanding
1706 of biology and uncover treatment paradigms for improving cardiometabolic health.
1707 DISCLOSURES
1708 SAS is a consultant, co-founder, and shareholder in Centaurus Therapeutics, LLC. The other
1711 All authors contributed to the concept and drafted manuscript equally. SAS and JLW edited and
1712 revised the manuscript. JLW prepared the figures. SMT compiled the table. WLH and SAS
1713 approved the final version. The authors received research support from the National Institutes of
78
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1714 Health (DK122001, DK131609, CA272529, DK116450, DK091317, and DK109894 to SAS;
1715 DK112826 and DK130296 to WLH; Ruth L. Kirschstein National Research Service Award
1717 1T32DK11096601 to SMT) the American Diabetes Association (to WLH), the Margolis
1718 Foundation (to WLH), and the American Heart Association (23PRE1020789, to SMT).
1719 ACKNOWLEDGEMENTS
1720 The chemical structures and Figures 1 and 2 were created in ChemDoodle. These figures, and
1722
1723
1724
1725
1726
79
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1727
1728
1729 Table 1
80
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
Whole Body Impaired skin barrier,
lethal(618, 619)
CerS3 Knockout
Overexpression Whole Body Alopecia(620)
(N/A)
CerS4 Knockout Ectodermal cells Hair loss(621)
81
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
lipolysis(632)
Intestinal Cells Nutrient uptake
regulation(633)
Smpd3 Knockout
Overexpression
(N/A)
82
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1730 REFERENCES
83
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1776 17. Hojjati MR, Li Z, and Jiang XC. Serine palmitoyl-CoA transferase (SPT) deficiency and
1777 sphingolipid levels in mice. Biochim Biophys Acta 1737: 44-51, 2005.
1778 18. Jiang XC, Goldberg IJ, and Park TS. Sphingolipids and cardiovascular diseases: lipoprotein
1779 metabolism, atherosclerosis and cardiomyopathy. Adv Exp Med Biol 721: 19-39, 2011.
1780 19. Park TS, Rosebury W, Kindt EK, Kowala MC, and Panek RL. Serine palmitoyltransferase inhibitor
1781 myriocin induces the regression of atherosclerotic plaques in hyperlipidemic ApoE-deficient mice.
1782 Pharmacol Res 58: 45-51, 2008.
1783 20. Russo SB, Baicu CF, Van Laer A, Geng T, Kasiganesan H, Zile MR, and Cowart LA. Ceramide
1784 synthase 5 mediates lipid-induced autophagy and hypertrophy in cardiomyocytes. J Clin Invest 122:
1785 3919-3930, 2012.
1786 21. Yang G, Badeanlou L, Bielawski J, Roberts AJ, Hannun YA, and Samad F. Central role of
1787 ceramide biosynthesis in body weight regulation, energy metabolism, and the metabolic syndrome. Am J
1788 Physiol Endocrinol Metab 297: E211-224, 2009.
1789 22. Park TS, Hu Y, Noh HL, Drosatos K, Okajima K, Buchanan J, Tuinei J, Homma S, Jiang XC, Abel
1790 ED, and Goldberg IJ. Ceramide is a cardiotoxin in lipotoxic cardiomyopathy. J Lipid Res 49: 2101-2112,
1791 2008.
1792 23. Polya JB, and Parsons RS. Free ceramide in blood and its relevance to atherosclerosis. I. The
1793 Medical journal of Australia 1: 873-879, 1973.
1794 24. Bismuth J, Lin P, Yao Q, and Chen C. Ceramide: a common pathway for atherosclerosis?
1795 Atherosclerosis 196: 497-504, 2008.
1796 25. Summers SA. Ceramides in insulin resistance and lipotoxicity. Prog Lipid Res 45: 42-72, 2006.
1797 26. Kolak M, Gertow J, Westerbacka J, Summers SA, Liska J, Franco-Cereceda A, Oresic M, Yki-
1798 Jarvinen H, Eriksson P, and Fisher RM. Expression of ceramide-metabolising enzymes in subcutaneous
1799 and intra-abdominal human adipose tissue. Lipids Health Dis 11: 115, 2012.
1800 27. Chavez JA, and Summers SA. A ceramide-centric view of insulin resistance. Cell Metab 15: 585-
1801 594, 2012.
1802 28. Bikman BT, and Summers SA. Ceramides as modulators of cellular and whole-body metabolism.
1803 J Clin Invest 121: 4222-4230, 2011.
1804 29. Bikman BT, and Summers SA. Sphingolipids and hepatic steatosis. Adv Exp Med Biol 721: 87-97,
1805 2011.
1806 30. Summers SA. Sphingolipids and insulin resistance: the five Ws. Curr Opin Lipidol 21: 128-135,
1807 2010.
1808 31. Holland WL, Knotts TA, Chavez JA, Wang LP, Hoehn KL, and Summers SA. Lipid mediators of
1809 insulin resistance. Nutr Rev 65: S39-46, 2007.
1810 32. Summers SA, and Nelson DH. A role for sphingolipids in producing the common features of type
1811 2 diabetes, metabolic syndrome X, and Cushing's syndrome. Diabetes 54: 591-602, 2005.
1812 33. Zhang QJ, Holland WL, Wilson L, Tanner JM, Kearns D, Cahoon JM, Pettey D, Losee J, Duncan B,
1813 Gale D, Kowalski CA, Deeter N, Nichols A, Deesing M, Arrant C, Ruan T, Boehme C, McCamey DR, Rou
1814 J, Ambal K, Narra KK, Summers SA, Abel ED, and Symons JD. Ceramide mediates vascular dysfunction in
1815 diet-induced obesity by PP2A-mediated dephosphorylation of the eNOS-Akt complex. Diabetes 61:
1816 1848-1859, 2012.
1817 34. Park TS, Panek RL, Rekhter MD, Mueller SB, Rosebury WS, Robertson A, Hanselman JC, Kindt
1818 E, Homan R, and Karathanasis SK. Modulation of lipoprotein metabolism by inhibition of sphingomyelin
1819 synthesis in ApoE knockout mice. Atherosclerosis 189: 264-272, 2006.
1820 35. Park TS, and Goldberg IJ. Sphingolipids, lipotoxic cardiomyopathy, and cardiac failure. Heart Fail
1821 Clin 8: 633-641, 2012.
1822 36. Lee SY, Kim JR, Hu Y, Khan R, Kim SJ, Bharadwaj KG, Davidson MM, Choi CS, Shin KO, Lee YM,
1823 Park WJ, Park IS, Jiang XC, Goldberg IJ, and Park TS. Cardiomyocyte specific deficiency of serine
84
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1824 palmitoyltransferase subunit 2 reduces ceramide but leads to cardiac dysfunction. J Biol Chem 287:
1825 18429-18439, 2012.
1826 37. Ji R, Akashi H, Drosatos K, Liao X, Jiang H, Kennel PJ, Brunjes DL, Castillero E, Zhang X, Deng LY,
1827 Homma S, George IJ, Takayama H, Naka Y, Goldberg IJ, and Schulze PC. Increased de novo ceramide
1828 synthesis and accumulation in failing myocardium. JCI Insight 2: 2017.
1829 38. Summers SA. Could Ceramides Become the New Cholesterol? Cell Metab 2017.
1830 39. Tippetts TS, Holland WL, and Summers SA. Cholesterol - the devil you know; ceramide - the
1831 devil you don't. Trends Pharmacol Sci 42: 1082-1095, 2021.
1832 40. Solving the Riddle of the Role of Sphingolipids in Cell Signaling. J Biol Chem 291: 11460-11461,
1833 2016.
1834 41. Hawthorne JN. A note on the life of J.L.W. Thudichum (1829-1901). Biochem Soc Trans 3: 591,
1835 1975.
1836 42. Hannun YA, Loomis CR, Merrill AH, Jr., and Bell RM. Sphingosine inhibition of protein kinase C
1837 activity and of phorbol dibutyrate binding in vitro and in human platelets. J Biol Chem 261: 12604-12609,
1838 1986.
1839 43. Wilson E, Olcott MC, Bell RM, Merrill AH, Jr., and Lambeth JD. Inhibition of the oxidative burst
1840 in human neutrophils by sphingoid long-chain bases. Role of protein kinase C in activation of the burst. J
1841 Biol Chem 261: 12616-12623, 1986.
1842 44. Merrill AH, Jr., Sereni AM, Stevens VL, Hannun YA, Bell RM, and Kinkade JM, Jr. Inhibition of
1843 phorbol ester-dependent differentiation of human promyelocytic leukemic (HL-60) cells by sphinganine
1844 and other long-chain bases. J Biol Chem 261: 12610-12615, 1986.
1845 45. Sbarra AJ, and Karnovsky ML. The biochemical basis of phagocytosis. I. Metabolic changes
1846 during the ingestion of particles by polymorphonuclear leukocytes. J Biol Chem 234: 1355-1362, 1959.
1847 46. Babior BM, Curnutte JT, and Kipnes RS. Biological defense mechanisms. Evidence for the
1848 participation of superoxide in bacterial killing by xanthine oxidase. J Lab Clin Med 85: 235-244, 1975.
1849 47. Babior BM. Oxidants from phagocytes: agents of defense and destruction. Blood 64: 959-966,
1850 1984.
1851 48. Rovera G, Santoli D, and Damsky C. Human promyelocytic leukemia cells in culture differentiate
1852 into macrophage-like cells when treated with a phorbol diester. Proc Natl Acad Sci U S A 76: 2779-2783,
1853 1979.
1854 49. Harris P, and Ralph P. Human leukemic models of myelomonocytic development: a review of
1855 the HL-60 and U937 cell lines. J Leukoc Biol 37: 407-422, 1985.
1856 50. Miyaura C, Abe E, Suda T, and Kuroki T. Alternative differentiation of human promyelocytic
1857 leukemia cells (HL-60) induced selectively by retinoic acid and 1 alpha,25-dihydroxyvitamin D3. Cancer
1858 Res 45: 4244-4248, 1985.
1859 51. Ghosh TK, Bian J, and Gill DL. Intracellular calcium release mediated by sphingosine derivatives
1860 generated in cells. Science 248: 1653-1656, 1990.
1861 52. Zhang H, Desai NN, Olivera A, Seki T, Brooker G, and Spiegel S. Sphingosine-1-phosphate, a
1862 novel lipid, involved in cellular proliferation. J Cell Biol 114: 155-167, 1991.
1863 53. Lee MJ, Van Brocklyn JR, Thangada S, Liu CH, Hand AR, Menzeleev R, Spiegel S, and Hla T.
1864 Sphingosine-1-phosphate as a ligand for the G protein-coupled receptor EDG-1. Science 279: 1552-1555,
1865 1998.
1866 54. Blaho VA, Galvani S, Engelbrecht E, Liu C, Swendeman SL, Kono M, Proia RL, Steinman L, Han
1867 MH, and Hla T. HDL-bound sphingosine-1-phosphate restrains lymphopoiesis and neuroinflammation.
1868 Nature 523: 342-346, 2015.
1869 55. Galvani S, Sanson M, Blaho VA, Swendeman SL, Obinata H, Conger H, Dahlback B, Kono M,
1870 Proia RL, Smith JD, and Hla T. HDL-bound sphingosine 1-phosphate acts as a biased agonist for the
1871 endothelial cell receptor S1P1 to limit vascular inflammation. Sci Signal 8: ra79, 2015.
85
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1872 56. Christensen PM, Liu CH, Swendeman SL, Obinata H, Qvortrup K, Nielsen LB, Hla T, Di Lorenzo
1873 A, and Christoffersen C. Impaired endothelial barrier function in apolipoprotein M-deficient mice is
1874 dependent on sphingosine-1-phosphate receptor 1. FASEB J 30: 2351-2359, 2016.
1875 57. Zhang W, Li Y, Li F, and Ling L. Sphingosine-1-phosphate receptor modulators in stroke
1876 treatment. J Neurochem 162: 390-403, 2022.
1877 58. Tolksdorf C, Moritz E, Wolf R, Meyer U, Marx S, Bien-Moller S, Garscha U, Jedlitschky G, and
1878 Rauch BH. Platelet-Derived S1P and Its Relevance for the Communication with Immune Cells in Multiple
1879 Human Diseases. Int J Mol Sci 23: 2022.
1880 59. Therond P, and Chapman MJ. Sphingosine-1-phosphate: metabolism, transport,
1881 atheroprotection and effect of statin treatment. Curr Opin Lipidol 33: 199-207, 2022.
1882 60. Spampinato SF, Sortino MA, and Salomone S. Sphingosine-1-phosphate and Sphingosine-1-
1883 phosphate receptors in the cardiovascular system: pharmacology and clinical implications. Adv
1884 Pharmacol 94: 95-139, 2022.
1885 61. Masuda-Kuroki K, and Di Nardo A. Sphingosine 1-Phosphate Signaling at the Skin Barrier
1886 Interface. Biology (Basel) 11: 2022.
1887 62. Bravo GA, Cedeno RR, Casadevall MP, and Ramio-Torrenta L. Sphingosine-1-Phosphate (S1P)
1888 and S1P Signaling Pathway Modulators, from Current Insights to Future Perspectives. Cells 11: 2022.
1889 63. Cartier A, and Hla T. Sphingosine 1-phosphate: Lipid signaling in pathology and therapy. Science
1890 366: 2019.
1891 64. Burg N, Salmon JE, and Hla T. Sphingosine 1-phosphate receptor-targeted therapeutics in
1892 rheumatic diseases. Nat Rev Rheumatol 18: 335-351, 2022.
1893 65. Janneh AH, and Ogretmen B. Targeting Sphingolipid Metabolism as a Therapeutic Strategy in
1894 Cancer Treatment. Cancers (Basel) 14: 2022.
1895 66. Di Pietro P, Izzo C, Abate AC, Iesu P, Rusciano MR, Venturini E, Visco V, Sommella E, Ciccarelli
1896 M, Carrizzo A, and Vecchione C. The Dark Side of Sphingolipids: Searching for Potential Cardiovascular
1897 Biomarkers. Biomolecules 13: 2023.
1898 67. Gaastra B, Zhang J, Tapper W, Bulters D, and Galea I. Sphingosine-1-phosphate Signalling in
1899 Aneurysmal Subarachnoid Haemorrhage: Basic Science to Clinical Translation. Transl Stroke Res 2023.
1900 68. Kleuser B, and Baumer W. Sphingosine 1-Phosphate as Essential Signaling Molecule in
1901 Inflammatory Skin Diseases. Int J Mol Sci 24: 2023.
1902 69. van Echten-Deckert G. The role of sphingosine 1-phosphate metabolism in brain health and
1903 disease. Pharmacol Ther 244: 108381, 2023.
1904 70. Wang N, Li JY, Zeng B, and Chen GL. Sphingosine-1-Phosphate Signaling in Cardiovascular
1905 Diseases. Biomolecules 13: 2023.
1906 71. Xiao J. Sphingosine 1-Phosphate Lyase in the Developing and Injured Nervous System: a
1907 Dichotomy? Mol Neurobiol 2023.
1908 72. Zhang F, and Lu Y. The Sphingosine 1-Phosphate Axis: an Emerging Therapeutic Opportunity for
1909 Endometriosis. Reprod Sci 30: 2040-2059, 2023.
1910 73. Okazaki T, Bielawska A, Bell RM, and Hannun YA. Role of ceramide as a lipid mediator of 1
1911 alpha,25-dihydroxyvitamin D3-induced HL-60 cell differentiation. J Biol Chem 265: 15823-15831, 1990.
1912 74. Igarashi Y, Kitamura K, Toyokuni T, Dean B, Fenderson B, Ogawass T, and Hakomori S. A
1913 specific enhancing effect of N,N-dimethylsphingosine on epidermal growth factor receptor
1914 autophosphorylation. Demonstration of its endogenous occurrence (and the virtual absence of
1915 unsubstituted sphingosine) in human epidermoid carcinoma A431 cells. J Biol Chem 265: 5385-5389,
1916 1990.
1917 75. Kolesnick RN. 1,2-Diacylglycerols but not phorbol esters stimulate sphingomyelin hydrolysis in
1918 GH3 pituitary cells. J Biol Chem 262: 16759-16762, 1987.
1919 76. Kolesnick RN. Sphingomyelin and derivatives as cellular signals. Prog Lipid Res 30: 1-38, 1991.
86
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1920 77. Kim MY, Linardic C, Obeid L, and Hannun Y. Identification of sphingomyelin turnover as an
1921 effector mechanism for the action of tumor necrosis factor alpha and gamma-interferon. Specific role in
1922 cell differentiation. J Biol Chem 266: 484-489, 1991.
1923 78. Hannun YA, and Obeid LM. Ceramide: an intracellular signal for apoptosis. Trends Biochem Sci
1924 20: 73-77, 1995.
1925 79. Kolesnick R. Ceramide: a novel second messenger. Trends Cell Biol 2: 232-236, 1992.
1926 80. Bose R, Verheij M, Haimovitz-Friedman A, Scotto K, Fuks Z, and Kolesnick R. Ceramide
1927 synthase mediates daunorubicin-induced apoptosis: an alternative mechanism for generating death
1928 signals. Cell 82: 405-414, 1995.
1929 81. Santana P, Pena LA, Haimovitz-Friedman A, Martin S, Green D, McLoughlin M, Cordon-Cardo
1930 C, Schuchman EH, Fuks Z, and Kolesnick R. Acid sphingomyelinase-deficient human lymphoblasts and
1931 mice are defective in radiation-induced apoptosis. Cell 86: 189-199, 1996.
1932 82. Hofmann K, and Dixit VM. Ceramide in apoptosis--does it really matter? Trends Biochem Sci 23:
1933 374-377, 1998.
1934 83. van Blitterswijk WJ, van der Luit AH, Veldman RJ, Verheij M, and Borst J. Ceramide: second
1935 messenger or modulator of membrane structure and dynamics? Biochem J 369: 199-211, 2003.
1936 84. Futerman AH, and Hannun YA. The complex life of simple sphingolipids. EMBO Rep 5: 777-782,
1937 2004.
1938 85. Merrill AH, Jr. Sphingolipid and glycosphingolipid metabolic pathways in the era of
1939 sphingolipidomics. Chem Rev 111: 6387-6422, 2011.
1940 86. Hannun YA, and Obeid LM. Sphingolipids and their metabolism in physiology and disease. Nat
1941 Rev Mol Cell Biol 19: 175-191, 2018.
1942 87. Merrill AH, Jr. De novo sphingolipid biosynthesis: a necessary, but dangerous, pathway. J Biol
1943 Chem 277: 25843-25846, 2002.
1944 88. Merrill AH, Jr., Schmelz EM, Dillehay DL, Spiegel S, Shayman JA, Schroeder JJ, Riley RT, Voss
1945 KA, and Wang E. Sphingolipids--the enigmatic lipid class: biochemistry, physiology, and pathophysiology.
1946 Toxicol Appl Pharmacol 142: 208-225, 1997.
1947 89. Han G, Gupta SD, Gable K, Niranjanakumari S, Moitra P, Eichler F, Brown RH, Jr., Harmon JM,
1948 and Dunn TM. Identification of small subunits of mammalian serine palmitoyltransferase that confer
1949 distinct acyl-CoA substrate specificities. Proc Natl Acad Sci U S A 106: 8186-8191, 2009.
1950 90. Eichler FS, Hornemann T, McCampbell A, Kuljis D, Penno A, Vardeh D, Tamrazian E, Garofalo K,
1951 Lee HJ, Kini L, Selig M, Frosch M, Gable K, von Eckardstein A, Woolf CJ, Guan G, Harmon JM, Dunn TM,
1952 and Brown RH, Jr. Overexpression of the wild-type SPT1 subunit lowers desoxysphingolipid levels and
1953 rescues the phenotype of HSAN1. J Neurosci 29: 14646-14651, 2009.
1954 91. Penno A, Reilly MM, Houlden H, Laura M, Rentsch K, Niederkofler V, Stoeckli ET, Nicholson G,
1955 Eichler F, Brown RH, Jr., von Eckardstein A, and Hornemann T. Hereditary sensory neuropathy type 1 is
1956 caused by the accumulation of two neurotoxic sphingolipids. J Biol Chem 285: 11178-11187, 2010.
1957 92. Lone MA, Santos T, Alecu I, Silva LC, and Hornemann T. 1-Deoxysphingolipids. Biochim Biophys
1958 Acta Mol Cell Biol Lipids 1864: 512-521, 2019.
1959 93. Bertea M, Rutti MF, Othman A, Marti-Jaun J, Hersberger M, von Eckardstein A, and
1960 Hornemann T. Deoxysphingoid bases as plasma markers in diabetes mellitus. Lipids Health Dis 9: 84,
1961 2010.
1962 94. Brozinick JT, Hawkins E, Hoang Bui H, Kuo MS, Tan B, Kievit P, and Grove K. Plasma
1963 sphingolipids are biomarkers of metabolic syndrome in non-human primates maintained on a Western-
1964 style diet. Int J Obes (Lond) 37: 1064-1070, 2013.
1965 95. Gai Z, Gui T, Alecu I, Lone MA, Hornemann T, Chen Q, Visentin M, Hiller C, Hausler S, and
1966 Kullak-Ublick GA. Farnesoid X receptor activation induces the degradation of hepatotoxic 1-
1967 deoxysphingolipids in non-alcoholic fatty liver disease. Liver Int 2019.
87
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
1968 96. Othman A, Bianchi R, Alecu I, Wei Y, Porretta-Serapiglia C, Lombardi R, Chiorazzi A, Meregalli
1969 C, Oggioni N, Cavaletti G, Lauria G, von Eckardstein A, and Hornemann T. Lowering plasma 1-
1970 deoxysphingolipids improves neuropathy in diabetic rats. Diabetes 64: 1035-1045, 2015.
1971 97. Othman A, Rutti MF, Ernst D, Saely CH, Rein P, Drexel H, Porretta-Serapiglia C, Lauria G,
1972 Bianchi R, von Eckardstein A, and Hornemann T. Plasma deoxysphingolipids: a novel class of biomarkers
1973 for the metabolic syndrome? Diabetologia 55: 421-431, 2012.
1974 98. Othman A, Saely CH, Muendlein A, Vonbank A, Drexel H, von Eckardstein A, and Hornemann
1975 T. Plasma 1-deoxysphingolipids are predictive biomarkers for type 2 diabetes mellitus. BMJ Open
1976 Diabetes Res Care 3: e000073, 2015.
1977 99. Handzlik MK, Gengatharan JM, Frizzi KE, McGregor GH, Martino C, Rahman G, Gonzalez A,
1978 Moreno AM, Green CR, Guernsey LS, Lin T, Tseng P, Ideguchi Y, Fallon RJ, Chaix A, Panda S, Mali P,
1979 Wallace M, Knight R, Gantner ML, Calcutt NA, and Metallo CM. Insulin-regulated serine and lipid
1980 metabolism drive peripheral neuropathy. Nature 614: 118-124, 2023.
1981 100. Cha HJ, He C, Zhao H, Dong Y, An IS, and An S. Intercellular and intracellular functions of
1982 ceramides and their metabolites in skin (Review). Int J Mol Med 38: 16-22, 2016.
1983 101. McCampbell A, Truong D, Broom DC, Allchorne A, Gable K, Cutler RG, Mattson MP, Woolf CJ,
1984 Frosch MP, Harmon JM, Dunn TM, and Brown RH, Jr. Mutant SPTLC1 dominantly inhibits serine
1985 palmitoyltransferase activity in vivo and confers an age-dependent neuropathy. Hum Mol Genet 14:
1986 3507-3521, 2005.
1987 102. Garofalo K, Penno A, Schmidt BP, Lee HJ, Frosch MP, von Eckardstein A, Brown RH,
1988 Hornemann T, and Eichler FS. Oral L-serine supplementation reduces production of neurotoxic
1989 deoxysphingolipids in mice and humans with hereditary sensory autonomic neuropathy type 1. J Clin
1990 Invest 121: 4735-4745, 2011.
1991 103. Hornemann T, Penno A, Rutti MF, Ernst D, Kivrak-Pfiffner F, Rohrer L, and von Eckardstein A.
1992 The SPTLC3 subunit of serine palmitoyltransferase generates short chain sphingoid bases. J Biol Chem
1993 284: 26322-26330, 2009.
1994 104. Russo SB, Tidhar R, Futerman AH, and Cowart LA. Myristate-derived d16:0 sphingolipids
1995 constitute a cardiac sphingolipid pool with distinct synthetic routes and functional properties. J Biol
1996 Chem 288: 13397-13409, 2013.
1997 105. Pewzner-Jung Y, Ben-Dor S, and Futerman AH. When do Lasses (longevity assurance genes)
1998 become CerS (ceramide synthases)?: Insights into the regulation of ceramide synthesis. J Biol Chem 281:
1999 25001-25005, 2006.
2000 106. Tidhar R, and Futerman AH. The complexity of sphingolipid biosynthesis in the endoplasmic
2001 reticulum. Biochim Biophys Acta 1833: 2511-2518, 2013.
2002 107. Levy M, and Futerman AH. Mammalian ceramide synthases. IUBMB Life 62: 347-356, 2010.
2003 108. Venkataraman K, Riebeling C, Bodennec J, Riezman H, Allegood JC, Sullards MC, Merrill AH, Jr.,
2004 and Futerman AH. Upstream of growth and differentiation factor 1 (uog1), a mammalian homolog of
2005 the yeast longevity assurance gene 1 (LAG1), regulates N-stearoyl-sphinganine (C18-(dihydro)ceramide)
2006 synthesis in a fumonisin B1-independent manner in mammalian cells. J Biol Chem 277: 35642-35649,
2007 2002.
2008 109. Laviad EL, Albee L, Pankova-Kholmyansky I, Epstein S, Park H, Merrill AH, Jr., and Futerman
2009 AH. Characterization of ceramide synthase 2: tissue distribution, substrate specificity, and inhibition by
2010 sphingosine 1-phosphate. J Biol Chem 283: 5677-5684, 2008.
2011 110. Mizutani Y, Kihara A, and Igarashi Y. LASS3 (longevity assurance homologue 3) is a mainly testis-
2012 specific (dihydro)ceramide synthase with relatively broad substrate specificity. Biochem J 398: 531-538,
2013 2006.
88
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2014 111. Riebeling C, Allegood JC, Wang E, Merrill AH, Jr., and Futerman AH. Two mammalian longevity
2015 assurance gene (LAG1) family members, trh1 and trh4, regulate dihydroceramide synthesis using
2016 different fatty acyl-CoA donors. J Biol Chem 278: 43452-43459, 2003.
2017 112. Mizutani Y, Kihara A, and Igarashi Y. Mammalian Lass6 and its related family members regulate
2018 synthesis of specific ceramides. Biochem J 390: 263-271, 2005.
2019 113. Harrer H, Laviad EL, Humpf HU, and Futerman AH. Identification of N-acyl-fumonisin B1 as new
2020 cytotoxic metabolites of fumonisin mycotoxins. Mol Nutr Food Res 57: 516-522, 2013.
2021 114. Kim HJ, Qiao Q, Toop HD, Morris JC, and Don AS. A fluorescent assay for ceramide synthase
2022 activity. J Lipid Res 53: 1701-1707, 2012.
2023 115. Lahiri S, Park H, Laviad EL, Lu X, Bittman R, and Futerman AH. Ceramide synthesis is modulated
2024 by the sphingosine analog FTY720 via a mixture of uncompetitive and noncompetitive inhibition in an
2025 Acyl-CoA chain length-dependent manner. J Biol Chem 284: 16090-16098, 2009.
2026 116. Zelnik ID, Rozman B, Rosenfeld-Gur E, Ben-Dor S, and Futerman AH. A Stroll Down the CerS
2027 Lane. Adv Exp Med Biol 1159: 49-63, 2019.
2028 117. Turpin SM, Nicholls HT, Willmes DM, Mourier A, Brodesser S, Wunderlich CM, Mauer J, Xu E,
2029 Hammerschmidt P, Bronneke HS, Trifunovic A, LoSasso G, Wunderlich FT, Kornfeld JW, Bluher M,
2030 Kronke M, and Bruning JC. Obesity-induced CerS6-dependent C16:0 ceramide production promotes
2031 weight gain and glucose intolerance. Cell Metab 20: 678-686, 2014.
2032 118. Raichur S, Wang ST, Chan PW, Li Y, Ching J, Chaurasia B, Dogra S, Ohman MK, Takeda K, Sugii
2033 S, Pewzner-Jung Y, Futerman AH, and Summers SA. CerS2 haploinsufficiency inhibits beta-oxidation and
2034 confers susceptibility to diet-induced steatohepatitis and insulin resistance. Cell Metab 20: 687-695,
2035 2014.
2036 119. Laviad EL, Kelly S, Merrill AH, Jr., and Futerman AH. Modulation of ceramide synthase activity
2037 via dimerization. J Biol Chem 287: 21025-21033, 2012.
2038 120. Pewzner-Jung Y, Brenner O, Braun S, Laviad EL, Ben-Dor S, Feldmesser E, Horn-Saban S,
2039 Amann-Zalcenstein D, Raanan C, Berkutzki T, Erez-Roman R, Ben-David O, Levy M, Holzman D, Park H,
2040 Nyska A, Merrill AH, Jr., and Futerman AH. A critical role for ceramide synthase 2 in liver homeostasis:
2041 II. insights into molecular changes leading to hepatopathy. J Biol Chem 285: 10911-10923, 2010.
2042 121. Mesicek J, Lee H, Feldman T, Jiang X, Skobeleva A, Berdyshev EV, Haimovitz-Friedman A, Fuks
2043 Z, and Kolesnick R. Ceramide synthases 2, 5, and 6 confer distinct roles in radiation-induced apoptosis in
2044 HeLa cells. Cell Signal 22: 1300-1307, 2010.
2045 122. Sridevi P, Alexander H, Laviad EL, Min J, Mesika A, Hannink M, Futerman AH, and Alexander S.
2046 Stress-induced ER to Golgi translocation of ceramide synthase 1 is dependent on proteasomal
2047 processing. Exp Cell Res 316: 78-91, 2010.
2048 123. Johnson KR, Johnson KY, Becker KP, Bielawski J, Mao C, and Obeid LM. Role of human
2049 sphingosine-1-phosphate phosphatase 1 in the regulation of intra- and extracellular sphingosine-1-
2050 phosphate levels and cell viability. J Biol Chem 278: 34541-34547, 2003.
2051 124. Maceyka M, Sankala H, Hait NC, Le Stunff H, Liu H, Toman R, Collier C, Zhang M, Satin LS,
2052 Merrill AH, Jr., Milstien S, and Spiegel S. SphK1 and SphK2, sphingosine kinase isoenzymes with
2053 opposing functions in sphingolipid metabolism. J Biol Chem 280: 37118-37129, 2005.
2054 125. Mandala SM, Thornton R, Galve-Roperh I, Poulton S, Peterson C, Olivera A, Bergstrom J, Kurtz
2055 MB, and Spiegel S. Molecular cloning and characterization of a lipid phosphohydrolase that degrades
2056 sphingosine-1- phosphate and induces cell death. Proc Natl Acad Sci U S A 97: 7859-7864, 2000.
2057 126. Osawa Y, Uchinami H, Bielawski J, Schwabe RF, Hannun YA, and Brenner DA. Roles for C16-
2058 ceramide and sphingosine 1-phosphate in regulating hepatocyte apoptosis in response to tumor
2059 necrosis factor-alpha. J Biol Chem 280: 27879-27887, 2005.
2060 127. Sociale M, Wulf AL, Breiden B, Klee K, Thielisch M, Eckardt F, Sellin J, Bulow MH, Lobbert S,
2061 Weinstock N, Voelzmann A, Schultze J, Sandhoff K, and Bauer R. Ceramide Synthase Schlank Is a
89
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2062 Transcriptional Regulator Adapting Gene Expression to Energy Requirements. Cell Rep 22: 967-978,
2063 2018.
2064 128. Cingolani F, Futerman AH, and Casas J. Ceramide synthases in biomedical research. Chem Phys
2065 Lipids 197: 25-32, 2016.
2066 129. Park JW, Park WJ, and Futerman AH. Ceramide synthases as potential targets for therapeutic
2067 intervention in human diseases. Biochim Biophys Acta 1841: 671-681, 2014.
2068 130. Senkal CE, Salama MF, Snider AJ, Allopenna JJ, Rana NA, Koller A, Hannun YA, and Obeid LM.
2069 Ceramide Is Metabolized to Acylceramide and Stored in Lipid Droplets. Cell Metab 25: 686-697, 2017.
2070 131. Rabionet M, Bayerle A, Marsching C, Jennemann R, Grone HJ, Yildiz Y, Wachten D, Shaw W,
2071 Shayman JA, and Sandhoff R. 1-O-acylceramides are natural components of human and mouse
2072 epidermis. J Lipid Res 54: 3312-3321, 2013.
2073 132. Sharma AX, Quittner-Strom EB, Lee Y, Johnson JA, Martin SA, Yu X, Li J, Lu J, Cai Z, Chen S,
2074 Wang MY, Zhang Y, Pearson MJ, Dorn AC, McDonald JG, Gordillo R, Yan H, Thai D, Wang ZV, Unger RH,
2075 and Holland WL. Glucagon Receptor Antagonism Improves Glucose Metabolism and Cardiac Function by
2076 Promoting AMP-Mediated Protein Kinase in Diabetic Mice. Cell Rep 22: 1760-1773, 2018.
2077 133. Shayman JA, and Abe A. 1-O-acylceramide synthase. Methods Enzymol 311: 105-117, 2000.
2078 134. Omae F, Miyazaki M, Enomoto A, Suzuki M, Suzuki Y, and Suzuki A. DES2 protein is responsible
2079 for phytoceramide biosynthesis in the mouse small intestine. Biochem J 379: 687-695, 2004.
2080 135. Ternes P, Franke S, Zahringer U, Sperling P, and Heinz E. Identification and characterization of a
2081 sphingolipid delta 4-desaturase family. J Biol Chem 277: 25512-25518, 2002.
2082 136. Poss AM, and Summers SA. Too Much of a Good Thing? An Evolutionary Theory to Explain the
2083 Role of Ceramides in NAFLD. Front Endocrinol (Lausanne) 11: 505, 2020.
2084 137. Futerman AH, Stieger B, Hubbard AL, and Pagano RE. Sphingomyelin synthesis in rat liver
2085 occurs predominantly at the cis and medial cisternae of the Golgi apparatus. J Biol Chem 265: 8650-
2086 8657, 1990.
2087 138. Astudillo L, Therville N, Colacios C, Segui B, Andrieu-Abadie N, and Levade T.
2088 Glucosylceramidases and malignancies in mammals. Biochimie 125: 267-280, 2016.
2089 139. Hanada K, Kumagai K, Yasuda S, Miura Y, Kawano M, Fukasawa M, and Nishijima M. Molecular
2090 machinery for non-vesicular trafficking of ceramide. Nature 426: 803-809, 2003.
2091 140. Chung LH, Liu D, Liu XT, and Qi Y. Ceramide Transfer Protein (CERT): An Overlooked Molecular
2092 Player in Cancer. Int J Mol Sci 22: 2021.
2093 141. Hanada K. Ceramide Transport from the Endoplasmic Reticulum to the Trans Golgi Region at
2094 Organelle Membrane Contact Sites. Adv Exp Med Biol 997: 69-81, 2017.
2095 142. Mencarelli C, Losen M, Hammels C, De Vry J, Hesselink MK, Steinbusch HW, De Baets MH, and
2096 Martinez-Martinez P. The ceramide transporter and the Goodpasture antigen binding protein: one
2097 protein--one function? J Neurochem 113: 1369-1386, 2010.
2098 143. Raya A, Revert F, Navarro S, and Saus J. Characterization of a novel type of serine/threonine
2099 kinase that specifically phosphorylates the human goodpasture antigen. J Biol Chem 274: 12642-12649,
2100 1999.
2101 144. Raya A, Revert-Ros F, Martinez-Martinez P, Navarro S, Rosello E, Vieites B, Granero F, Forteza
2102 J, and Saus J. Goodpasture antigen-binding protein, the kinase that phosphorylates the goodpasture
2103 antigen, is an alternatively spliced variant implicated in autoimmune pathogenesis. J Biol Chem 275:
2104 40392-40399, 2000.
2105 145. Wang X, Rao RP, Kosakowska-Cholody T, Masood MA, Southon E, Zhang H, Berthet C,
2106 Nagashim K, Veenstra TK, Tessarollo L, Acharya U, and Acharya JK. Mitochondrial degeneration and not
2107 apoptosis is the primary cause of embryonic lethality in ceramide transfer protein mutant mice. J Cell
2108 Biol 184: 143-158, 2009.
90
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2109 146. Soccio RE, and Breslow JL. StAR-related lipid transfer (START) proteins: mediators of
2110 intracellular lipid metabolism. J Biol Chem 278: 22183-22186, 2003.
2111 147. Kumagai K, Yasuda S, Okemoto K, Nishijima M, Kobayashi S, and Hanada K. CERT mediates
2112 intermembrane transfer of various molecular species of ceramides. J Biol Chem 280: 6488-6495, 2005.
2113 148. Chandran S, and Machamer CE. Inactivation of ceramide transfer protein during pro-apoptotic
2114 stress by Golgi disassembly and caspase cleavage. Biochem J 442: 391-401, 2012.
2115 149. Heering J, Weis N, Holeiter M, Neugart F, Staebler A, Fehm TN, Bischoff A, Schiller J, Duss S,
2116 Schmid S, Korte T, Herrmann A, and Olayioye MA. Loss of the ceramide transfer protein augments EGF
2117 receptor signaling in breast cancer. Cancer Res 72: 2855-2866, 2012.
2118 150. Derre I, Swiss R, and Agaisse H. The lipid transfer protein CERT interacts with the Chlamydia
2119 inclusion protein IncD and participates to ER-Chlamydia inclusion membrane contact sites. PLoS Pathog
2120 7: e1002092, 2011.
2121 151. Amako Y, Syed GH, and Siddiqui A. Protein kinase D negatively regulates hepatitis C virus
2122 secretion through phosphorylation of oxysterol-binding protein and ceramide transfer protein. J Biol
2123 Chem 286: 11265-11274, 2011.
2124 152. Olayioye MA, and Hausser A. Integration of non-vesicular and vesicular transport processes at
2125 the Golgi complex by the PKD-CERT network. Biochim Biophys Acta 1821: 1096-1103, 2012.
2126 153. Kudo N, Kumagai K, Tomishige N, Yamaji T, Wakatsuki S, Nishijima M, Hanada K, and Kato R.
2127 Structural basis for specific lipid recognition by CERT responsible for nonvesicular trafficking of
2128 ceramide. Proc Natl Acad Sci U S A 105: 488-493, 2008.
2129 154. Peretti D, Dahan N, Shimoni E, Hirschberg K, and Lev S. Coordinated lipid transfer between the
2130 endoplasmic reticulum and the Golgi complex requires the VAP proteins and is essential for Golgi-
2131 mediated transport. Mol Biol Cell 19: 3871-3884, 2008.
2132 155. Tuuf J, Kjellberg MA, Molotkovsky JG, Hanada K, and Mattjus P. The intermembrane ceramide
2133 transport catalyzed by CERT is sensitive to the lipid environment. Biochim Biophys Acta 1808: 229-235,
2134 2011.
2135 156. Bornancin F. Ceramide kinase: the first decade. Cell Signal 23: 999-1008, 2011.
2136 157. Zhang Y, Zhang X, Lu M, and Zou X. Ceramide-1-phosphate and its transfer proteins in
2137 eukaryotes. Chem Phys Lipids 240: 105135, 2021.
2138 158. Ouro A, Arana L, Gangoiti P, Rivera IG, Ordonez M, Trueba M, Lankalapalli RS, Bittman R, and
2139 Gomez-Munoz A. Ceramide 1-phosphate stimulates glucose uptake in macrophages. Cell Signal 25: 786-
2140 795, 2013.
2141 159. Futerman AH, and Pagano RE. Determination of the intracellular sites and topology of
2142 glucosylceramide synthesis in rat liver. Biochem J 280 ( Pt 2): 295-302, 1991.
2143 160. Nylund M, Kjellberg MA, Molotkovsky JG, Byun HS, Bittman R, and Mattjus P. Molecular
2144 features of phospholipids that affect glycolipid transfer protein-mediated galactosylceramide transfer
2145 between vesicles. Biochim Biophys Acta 1758: 807-812, 2006.
2146 161. D'Angelo G, Polishchuk E, Di Tullio G, Santoro M, Di Campli A, Godi A, West G, Bielawski J,
2147 Chuang CC, van der Spoel AC, Platt FM, Hannun YA, Polishchuk R, Mattjus P, and De Matteis MA.
2148 Glycosphingolipid synthesis requires FAPP2 transfer of glucosylceramide. Nature 449: 62-67, 2007.
2149 162. Hakomori S. Traveling for the glycosphingolipid path. Glycoconj J 17: 627-647, 2000.
2150 163. Ichikawa S, and Hirabayashi Y. Glucosylceramide synthase and glycosphingolipid synthesis.
2151 Trends Cell Biol 8: 198-202, 1998.
2152 164. Marchesini N, and Hannun YA. Acid and neutral sphingomyelinases: roles and mechanisms of
2153 regulation. Biochem Cell Biol 82: 27-44, 2004.
2154 165. Schuchman EH, Suchi M, Takahashi T, Sandhoff K, and Desnick RJ. Human acid
2155 sphingomyelinase. Isolation, nucleotide sequence and expression of the full-length and alternatively
2156 spliced cDNAs. J Biol Chem 266: 8531-8539, 1991.
91
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2157 166. Kitatani K, Sheldon K, Rajagopalan V, Anelli V, Jenkins RW, Sun Y, Grabowski GA, Obeid LM,
2158 and Hannun YA. Involvement of acid beta-glucosidase 1 in the salvage pathway of ceramide formation. J
2159 Biol Chem 284: 12972-12978, 2009.
2160 167. Inokuchi J. GM3 and diabetes. Glycoconj J 31: 193-197, 2014.
2161 168. Lipina C, and Hundal HS. Ganglioside GM3 as a gatekeeper of obesity-associated insulin
2162 resistance: Evidence and mechanisms. FEBS Lett 589: 3221-3227, 2015.
2163 169. Vukovic I, Bozic J, Markotic A, Ljubicic S, and Ticinovic Kurir T. The missing link - likely
2164 pathogenetic role of GM3 and other gangliosides in the development of diabetic nephropathy. Kidney
2165 Blood Press Res 40: 306-314, 2015.
2166 170. Claus RA, Dorer MJ, Bunck AC, and Deigner HP. Inhibition of sphingomyelin hydrolysis: targeting
2167 the lipid mediator ceramide as a key regulator of cellular fate. Curr Med Chem 16: 1978-2000, 2009.
2168 171. Bienias K, Fiedorowicz A, Sadowska A, Prokopiuk S, and Car H. Regulation of sphingomyelin
2169 metabolism. Pharmacol Rep 68: 570-581, 2016.
2170 172. Marathe S, Schissel SL, Yellin MJ, Beatini N, Mintzer R, Williams KJ, and Tabas I. Human
2171 vascular endothelial cells are a rich and regulatable source of secretory sphingomyelinase. Implications
2172 for early atherogenesis and ceramide-mediated cell signaling. J Biol Chem 273: 4081-4088, 1998.
2173 173. Schissel SL, Jiang X, Tweedie-Hardman J, Jeong T, Camejo EH, Najib J, Rapp JH, Williams KJ, and
2174 Tabas I. Secretory sphingomyelinase, a product of the acid sphingomyelinase gene, can hydrolyze
2175 atherogenic lipoproteins at neutral pH. Implications for atherosclerotic lesion development. J Biol Chem
2176 273: 2738-2746, 1998.
2177 174. Schissel SL, Schuchman EH, Williams KJ, and Tabas I. Zn2+-stimulated sphingomyelinase is
2178 secreted by many cell types and is a product of the acid sphingomyelinase gene. J Biol Chem 271: 18431-
2179 18436, 1996.
2180 175. Ferlinz K, Hurwitz R, Moczall H, Lansmann S, Schuchman EH, and Sandhoff K. Functional
2181 characterization of the N-glycosylation sites of human acid sphingomyelinase by site-directed
2182 mutagenesis. Eur J Biochem 243: 511-517, 1997.
2183 176. Levade T, Salvayre R, and Douste-Blazy L. Sphingomyelinases and Niemann-Pick disease. J Clin
2184 Chem Clin Biochem 24: 205-220, 1986.
2185 177. Opreanu M, Tikhonenko M, Bozack S, Lydic TA, Reid GE, McSorley KM, Sochacki A, Perez GI,
2186 Esselman WJ, Kern T, Kolesnick R, Grant MB, and Busik JV. The unconventional role of acid
2187 sphingomyelinase in regulation of retinal microangiopathy in diabetic human and animal models.
2188 Diabetes 60: 2370-2378, 2011.
2189 178. Spence MW, Byers DM, Palmer FB, and Cook HW. A new Zn2+-stimulated sphingomyelinase in
2190 fetal bovine serum. J Biol Chem 264: 5358-5363, 1989.
2191 179. Hofmann K, Tomiuk S, Wolff G, and Stoffel W. Cloning and characterization of the mammalian
2192 brain-specific, Mg2+-dependent neutral sphingomyelinase. Proc Natl Acad Sci U S A 97: 5895-5900,
2193 2000.
2194 180. Tomiuk S, Hofmann K, Nix M, Zumbansen M, and Stoffel W. Cloned mammalian neutral
2195 sphingomyelinase: functions in sphingolipid signaling? Proc Natl Acad Sci U S A 95: 3638-3643, 1998.
2196 181. Tani M, and Hannun YA. Neutral sphingomyelinase 2 is palmitoylated on multiple cysteine
2197 residues. Role of palmitoylation in subcellular localization. J Biol Chem 282: 10047-10056, 2007.
2198 182. Tomiuk S, Zumbansen M, and Stoffel W. Characterization and subcellular localization of murine
2199 and human magnesium-dependent neutral sphingomyelinase. J Biol Chem 275: 5710-5717, 2000.
2200 183. Mizutani Y, Tamiya-Koizumi K, Nakamura N, Kobayashi M, Hirabayashi Y, and Yoshida S.
2201 Nuclear localization of neutral sphingomyelinase 1: biochemical and immunocytochemical analyses. J
2202 Cell Sci 114: 3727-3736, 2001.
92
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2203 184. Krut O, Wiegmann K, Kashkar H, Yazdanpanah B, and Kronke M. Novel tumor necrosis factor-
2204 responsive mammalian neutral sphingomyelinase-3 is a C-tail-anchored protein. J Biol Chem 281: 13784-
2205 13793, 2006.
2206 185. Rajagopalan V, Canals D, Luberto C, Snider J, Voelkel-Johnson C, Obeid LM, and Hannun YA.
2207 Critical determinants of mitochondria-associated neutral sphingomyelinase (MA-nSMase) for
2208 mitochondrial localization. Biochim Biophys Acta 1850: 628-639, 2015.
2209 186. Wu BX, Rajagopalan V, Roddy PL, Clarke CJ, and Hannun YA. Identification and characterization
2210 of murine mitochondria-associated neutral sphingomyelinase (MA-nSMase), the mammalian
2211 sphingomyelin phosphodiesterase 5. J Biol Chem 285: 17993-18002, 2010.
2212 187. Duan RD, Bergman T, Xu N, Wu J, Cheng Y, Duan J, Nelander S, Palmberg C, and Nilsson A.
2213 Identification of human intestinal alkaline sphingomyelinase as a novel ecto-enzyme related to the
2214 nucleotide phosphodiesterase family. J Biol Chem 278: 38528-38536, 2003.
2215 188. Duan RD, Cheng Y, Hansen G, Hertervig E, Liu JJ, Syk I, Sjostrom H, and Nilsson A. Purification,
2216 localization, and expression of human intestinal alkaline sphingomyelinase. J Lipid Res 44: 1241-1250,
2217 2003.
2218 189. Coant N, Sakamoto W, Mao C, and Hannun YA. Ceramidases, roles in sphingolipid metabolism
2219 and in health and disease. Adv Biol Regul 63: 122-131, 2017.
2220 190. Vasiliauskaité-Brooks I, Sounier R, Rochaix P, Bellot G, Fortier M, Hoh F, De Colibus L, Bechara
2221 C, Saied EM, Arenz C, Leyrat C, and Granier S. Structural insights into adiponectin receptors suggest
2222 ceramidase activity. Nature 544: 120-123, 2017.
2223 191. Gatt S. Enzymatic hydrolysis of sphingolipids. I. Hydrolysis and synthesis of ceramides by an
2224 enzyme from rat brain. J Biol Chem 241: 3724-3730, 1966.
2225 192. Sugita M, Dulaney JT, and Moser HW. Ceramidase deficiency in Farber's disease
2226 (lipogranulomatosis). Science 178: 1100-1102, 1972.
2227 193. Li CM, Park JH, He X, Levy B, Chen F, Arai K, Adler DA, Disteche CM, Koch J, Sandhoff K, and
2228 Schuchman EH. The human acid ceramidase gene (ASAH): structure, chromosomal location, mutation
2229 analysis, and expression. Genomics 62: 223-231, 1999.
2230 194. Mao C, and Obeid LM. Ceramidases: regulators of cellular responses mediated by ceramide,
2231 sphingosine, and sphingosine-1-phosphate. Biochim Biophys Acta 1781: 424-434, 2008.
2232 195. Okino N, He X, Gatt S, Sandhoff K, Ito M, and Schuchman EH. The reverse activity of human
2233 acid ceramidase. J Biol Chem 278: 29948-29953, 2003.
2234 196. Eliyahu E, Park JH, Shtraizent N, He X, and Schuchman EH. Acid ceramidase is a novel factor
2235 required for early embryo survival. FASEB J 21: 1403-1409, 2007.
2236 197. Chaurasia B, Ying L, Talbot CL, Maschek JA, Cox J, Schuchman EH, Hirabayashi Y, Holland WL,
2237 and Summers SA. Ceramides are necessary and sufficient for diet-induced impairment of thermogenic
2238 adipocytes. Mol Metab 45: 101145, 2021.
2239 198. Kono M, Dreier JL, Ellis JM, Allende ML, Kalkofen DN, Sanders KM, Bielawski J, Bielawska A,
2240 Hannun YA, and Proia RL. Neutral ceramidase encoded by the Asah2 gene is essential for the intestinal
2241 degradation of sphingolipids. J Biol Chem 281: 7324-7331, 2006.
2242 199. Franzen R, Pfeilschifter J, and Huwiler A. Nitric oxide induces neutral ceramidase degradation
2243 by the ubiquitin/proteasome complex in renal mesangial cell cultures. FEBS Lett 532: 441-444, 2002.
2244 200. Franzen R, Fabbro D, Aschrafi A, Pfeilschifter J, and Huwiler A. Nitric oxide induces degradation
2245 of the neutral ceramidase in rat renal mesangial cells and is counterregulated by protein kinase C. J Biol
2246 Chem 277: 46184-46190, 2002.
2247 201. Mitsutake S, Tani M, Okino N, Mori K, Ichinose S, Omori A, Iida H, Nakamura T, and Ito M.
2248 Purification, characterization, molecular cloning, and subcellular distribution of neutral ceramidase of
2249 rat kidney. J Biol Chem 276: 26249-26259, 2001.
93
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2250 202. Franzen R, Pautz A, Brautigam L, Geisslinger G, Pfeilschifter J, and Huwiler A. Interleukin-1beta
2251 induces chronic activation and de novo synthesis of neutral ceramidase in renal mesangial cells. J Biol
2252 Chem 276: 35382-35389, 2001.
2253 203. Liakath-Ali K, Vancollie VE, Lelliott CJ, Speak AO, Lafont D, Protheroe HJ, Ingvorsen C, Galli A,
2254 Green A, Gleeson D, Ryder E, Glover L, Vizcay-Barrena G, Karp NA, Arends MJ, Brenn T, Spiegel S,
2255 Adams DJ, Watt FM, and van der Weyden L. Alkaline ceramidase 1 is essential for mammalian skin
2256 homeostasis and regulating whole-body energy expenditure. J Pathol 239: 374-383, 2016.
2257 204. Sun W, Xu R, Hu W, Jin J, Crellin HA, Bielawski J, Szulc ZM, Thiers BH, Obeid LM, and Mao C.
2258 Upregulation of the human alkaline ceramidase 1 and acid ceramidase mediates calcium-induced
2259 differentiation of epidermal keratinocytes. J Invest Dermatol 128: 389-397, 2008.
2260 205. Huang Z, Huang G, Li Q, and Jin J. p38 mitogen-activated protein kinase/activator protein-1
2261 involved in serum deprivation-induced human alkaline ceramidase 2 upregulation. Biomed Rep 3: 225-
2262 229, 2015.
2263 206. Xu R, Jin J, Hu W, Sun W, Bielawski J, Szulc Z, Taha T, Obeid LM, and Mao C. Golgi alkaline
2264 ceramidase regulates cell proliferation and survival by controlling levels of sphingosine and S1P. FASEB J
2265 20: 1813-1825, 2006.
2266 207. Xu R, Wang K, Mileva I, Hannun YA, Obeid LM, and Mao C. Alkaline ceramidase 2 and its
2267 bioactive product sphingosine are novel regulators of the DNA damage response. Oncotarget 7: 18440-
2268 18457, 2016.
2269 208. Mao C, Xu R, Szulc ZM, Bielawska A, Galadari SH, and Obeid LM. Cloning and characterization
2270 of a novel human alkaline ceramidase. A mammalian enzyme that hydrolyzes phytoceramide. J Biol
2271 Chem 276: 26577-26588, 2001.
2272 209. Hu W, Xu R, Sun W, Szulc ZM, Bielawski J, Obeid LM, and Mao C. Alkaline ceramidase 3 (ACER3)
2273 hydrolyzes unsaturated long-chain ceramides, and its down-regulation inhibits both cell proliferation
2274 and apoptosis. J Biol Chem 285: 7964-7976, 2010.
2275 210. Wang K, Xu R, Schrandt J, Shah P, Gong YZ, Preston C, Wang L, Yi JK, Lin CL, Sun W,
2276 Spyropoulos DD, Rhee S, Li M, Zhou J, Ge S, Zhang G, Snider AJ, Hannun YA, Obeid LM, and Mao C.
2277 Alkaline Ceramidase 3 Deficiency Results in Purkinje Cell Degeneration and Cerebellar Ataxia Due to
2278 Dyshomeostasis of Sphingolipids in the Brain. PLoS Genet 11: e1005591, 2015.
2279 211. Ding G, Sonoda H, Yu H, Kajimoto T, Goparaju SK, Jahangeer S, Okada T, and Nakamura S.
2280 Protein kinase D-mediated phosphorylation and nuclear export of sphingosine kinase 2. J Biol Chem 282:
2281 27493-27502, 2007.
2282 212. Igarashi N, Okada T, Hayashi S, Fujita T, Jahangeer S, and Nakamura S. Sphingosine kinase 2 is a
2283 nuclear protein and inhibits DNA synthesis. J Biol Chem 278: 46832-46839, 2003.
2284 213. Don AS, and Rosen H. A lipid binding domain in sphingosine kinase 2. Biochem Biophys Res
2285 Commun 380: 87-92, 2009.
2286 214. Kohama T, Olivera A, Edsall L, Nagiec MM, Dickson R, and Spiegel S. Molecular cloning and
2287 functional characterization of murine sphingosine kinase. J Biol Chem 273: 23722-23728, 1998.
2288 215. Neubauer HA, and Pitson SM. Roles, regulation and inhibitors of sphingosine kinase 2. FEBS J
2289 280: 5317-5336, 2013.
2290 216. Spiegel S, Maczis MA, Maceyka M, and Milstien S. New insights into functions of the
2291 sphingosine-1-phosphate transporter SPNS2. J Lipid Res 60: 484-489, 2019.
2292 217. Maceyka M, and Spiegel S. Sphingolipid metabolites in inflammatory disease. Nature 510: 58-
2293 67, 2014.
2294 218. Le Stunff H, Giussani P, Maceyka M, Lepine S, Milstien S, and Spiegel S. Recycling of
2295 sphingosine is regulated by the concerted actions of sphingosine-1-phosphate phosphohydrolase 1 and
2296 sphingosine kinase 2. J Biol Chem 282: 34372-34380, 2007.
94
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2297 219. Ogawa C, Kihara A, Gokoh M, and Igarashi Y. Identification and characterization of a novel
2298 human sphingosine-1-phosphate phosphohydrolase, hSPP2. J Biol Chem 278: 1268-1272, 2003.
2299 220. Nakahara K, Ohkuni A, Kitamura T, Abe K, Naganuma T, Ohno Y, Zoeller RA, and Kihara A. The
2300 Sjogren-Larsson syndrome gene encodes a hexadecenal dehydrogenase of the sphingosine 1-phosphate
2301 degradation pathway. Mol Cell 46: 461-471, 2012.
2302 221. Sasset L, Zhang Y, Dunn TM, and Di Lorenzo A. Sphingolipid De Novo Biosynthesis: A Rheostat
2303 of Cardiovascular Homeostasis. Trends Endocrinol Metab 27: 807-819, 2016.
2304 222. Breslow DK, Collins SR, Bodenmiller B, Aebersold R, Simons K, Shevchenko A, Ejsing CS, and
2305 Weissman JS. Orm family proteins mediate sphingolipid homeostasis. Nature 463: 1048-1053, 2010.
2306 223. Roelants FM, Breslow DK, Muir A, Weissman JS, and Thorner J. Protein kinase Ypk1
2307 phosphorylates regulatory proteins Orm1 and Orm2 to control sphingolipid homeostasis in
2308 Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 108: 19222-19227, 2011.
2309 224. Sun Y, Miao Y, Yamane Y, Zhang C, Shokat KM, Takematsu H, Kozutsumi Y, and Drubin DG.
2310 Orm protein phosphoregulation mediates transient sphingolipid biosynthesis response to heat stress via
2311 the Pkh-Ypk and Cdc55-PP2A pathways. Mol Biol Cell 23: 2388-2398, 2012.
2312 225. Chauhan N, Han G, Somashekarappa N, Gable K, Dunn T, and Kohlwein SD. Regulation of
2313 Sphingolipid Biosynthesis by the Morphogenesis Checkpoint Kinase Swe1. J Biol Chem 291: 2524-2534,
2314 2016.
2315 226. Hjelmqvist L, Tuson M, Marfany G, Herrero E, Balcells S, and Gonzalez-Duarte R. ORMDL
2316 proteins are a conserved new family of endoplasmic reticulum membrane proteins. Genome Biol 3:
2317 RESEARCH0027, 2002.
2318 227. Sasset L, Chowdhury KH, Manzo OL, Rubinelli L, Konrad C, Maschek JA, Manfredi G, Holland
2319 WL, and Di Lorenzo A. Sphingosine-1-phosphate controls endothelial sphingolipid homeostasis via
2320 ORMDL. EMBO Rep 24: e54689, 2023.
2321 228. Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, von Kriegsheim A, Hebestreit
2322 HF, Mukherji M, Schofield CJ, Maxwell PH, Pugh CW, and Ratcliffe PJ. Targeting of HIF-alpha to the von
2323 Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292: 468-472, 2001.
2324 229. Yu M, Lun J, Zhang H, Zhu L, Zhang G, and Fang J. The non-canonical functions of HIF prolyl
2325 hydroxylases and their dual roles in cancer. Int J Biochem Cell Biol 135: 105982, 2021.
2326 230. Siow D, Sunkara M, Morris A, and Wattenberg B. Regulation of de novo sphingolipid
2327 biosynthesis by the ORMDL proteins and sphingosine kinase-1. Adv Biol Regul 57: 42-54, 2015.
2328 231. Davis DL, Gable K, Suemitsu J, Dunn TM, and Wattenberg BW. The ORMDL/Orm-serine
2329 palmitoyltransferase (SPT) complex is directly regulated by ceramide: Reconstitution of SPT regulation in
2330 isolated membranes. J Biol Chem 294: 5146-5156, 2019.
2331 232. Mohassel P, Donkervoort S, Lone MA, Nalls M, Gable K, Gupta SD, Foley AR, Hu Y, Saute JAM,
2332 Moreira AL, Kok F, Introna A, Logroscino G, Grunseich C, Nickolls AR, Pourshafie N, Neuhaus SB, Saade
2333 D, Gangfuss A, Kolbel H, Piccus Z, Le Pichon CE, Fiorillo C, Ly CV, Topf A, Brady L, Specht S, Zidell A,
2334 Pedro H, Mittelmann E, Thomas FP, Chao KR, Konersman CG, Cho MT, Brandt T, Straub V, Connolly
2335 AM, Schara U, Roos A, Tarnopolsky M, Hoke A, Brown RH, Lee CH, Hornemann T, Dunn TM, and
2336 Bonnemann CG. Childhood amyotrophic lateral sclerosis caused by excess sphingolipid synthesis. Nat
2337 Med 27: 1197-1204, 2021.
2338 233. Sasset L, Manzo OL, Zhang Y, Marino A, Rubinelli L, Riemma MA, Chalasani MLS, Dasoveanu
2339 DC, Roviezzo F, Jankauskas SS, Santulli G, Bucci MR, Lu TT, and Di Lorenzo A. Nogo-A reduces ceramide
2340 de novo biosynthesis to protect from heart failure. Cardiovasc Res 2022.
2341 234. Sasset L, and Di Lorenzo A. Sphingolipid Metabolism and Signaling in Endothelial Cell Functions.
2342 Adv Exp Med Biol 1372: 87-117, 2022.
95
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2343 235. Zhang Y, Huang Y, Cantalupo A, Azevedo PS, Siragusa M, Bielawski J, Giordano FJ, and Di
2344 Lorenzo A. Endothelial Nogo-B regulates sphingolipid biosynthesis to promote pathological cardiac
2345 hypertrophy during chronic pressure overload. JCI Insight 1: 2016.
2346 236. Sasset L, Manzo OL, Zhang Y, Marino A, Rubinelli L, Riemma MA, Chalasani MLS, Dasoveanu
2347 DC, Roviezzo F, Jankauskas SS, Santulli G, Bucci MR, Lu TT, and Di Lorenzo A. Nogo-A reduces ceramide
2348 de novo biosynthesis to protect from heart failure. Cardiovasc Res 119: 506-519, 2023.
2349 237. Pathak GP, Shah R, Kennedy BE, Murphy JP, Clements D, Konda P, Giacomantonio M, Xu Z,
2350 Schlaepfer IR, and Gujar S. RTN4 Knockdown Dysregulates the AKT Pathway, Destabilizes the
2351 Cytoskeleton, and Enhances Paclitaxel-Induced Cytotoxicity in Cancers. Mol Ther 26: 2019-2033, 2018.
2352 238. Chavez JA, and Summers SA. Characterizing the effects of saturated fatty acids on insulin
2353 signaling and ceramide and diacylglycerol accumulation in 3T3-L1 adipocytes and C2C12 myotubes. Arch
2354 Biochem Biophys 419: 101-109, 2003.
2355 239. Stockli J, Fisher-Wellman KH, Chaudhuri R, Zeng XY, Fazakerley DJ, Meoli CC, Thomas KC,
2356 Hoffman NJ, Mangiafico SP, Xirouchaki CE, Yang CH, Ilkayeva O, Wong K, Cooney GJ, Andrikopoulos S,
2357 Muoio DM, and James DE. Metabolomic analysis of insulin resistance across different mouse strains and
2358 diets. J Biol Chem 292: 19135-19145, 2017.
2359 240. Luukkonen PK, Sadevirta S, Zhou Y, Kayser B, Ali A, Ahonen L, Lallukka S, Pelloux V, Gaggini M,
2360 Jian C, Hakkarainen A, Lundbom N, Gylling H, Salonen A, Oresic M, Hyotylainen T, Orho-Melander M,
2361 Rissanen A, Gastaldelli A, Clement K, Hodson L, and Yki-Jarvinen H. Saturated Fat Is More Metabolically
2362 Harmful for the Human Liver Than Unsaturated Fat or Simple Sugars. Diabetes Care 41: 1732-1739,
2363 2018.
2364 241. Kien CL, Bunn JY, Poynter ME, Stevens R, Bain J, Ikayeva O, Fukagawa NK, Champagne CM,
2365 Crain KI, Koves TR, and Muoio DM. A lipidomics analysis of the relationship between dietary fatty acid
2366 composition and insulin sensitivity in young adults. Diabetes 62: 1054-1063, 2013.
2367 242. Huang H, Kasumov T, Gatmaitan P, Heneghan HM, Kashyap SR, Schauer PR, Brethauer SA, and
2368 Kirwan JP. Gastric bypass surgery reduces plasma ceramide subspecies and improves insulin sensitivity
2369 in severely obese patients. Obesity (Silver Spring) 19: 2235-2240, 2011.
2370 243. Heneghan HM, Huang H, Kashyap SR, Gornik HL, McCullough AJ, Schauer PR, Brethauer SA,
2371 Kirwan JP, and Kasumov T. Reduced cardiovascular risk after bariatric surgery is linked to plasma
2372 ceramides, apolipoprotein-B100, and ApoB100/A1 ratio. Surg Obes Relat Dis 9: 100-107, 2013.
2373 244. Özer H, Aslan İ, Oruç MT, Çöpelci Y, Afşar E, Kaya S, and Aslan M. Early postoperative changes
2374 of sphingomyelins and ceramides after laparoscopic sleeve gastrectomy. Lipids Health Dis 17: 269, 2018.
2375 245. Poss AM, Krick B, Maschek JA, Haaland B, Cox JE, Karra P, Ibele AR, Hunt SC, Adams TD,
2376 Holland WL, Playdon MC, and Summers SA. Following Roux-en-Y gastric bypass surgery, serum
2377 ceramides demarcate patients that will fail to achieve normoglycemia and diabetes remission. Med (N Y)
2378 3: 452-467 e454, 2022.
2379 246. Muthusamy T, Cordes T, Handzlik MK, You L, Lim EW, Gengatharan J, Pinto AFM, Badur MG,
2380 Kolar MJ, Wallace M, Saghatelian A, and Metallo CM. Serine restriction alters sphingolipid diversity to
2381 constrain tumour growth. Nature 586: 790-795, 2020.
2382 247. Miller RA, Shi Y, Lu W, Pirman DA, Jatkar A, Blatnik M, Wu H, Cardenas C, Wan M, Foskett JK,
2383 Park JO, Zhang Y, Holland WL, Rabinowitz JD, and Birnbaum MJ. Targeting hepatic glutaminase activity
2384 to ameliorate hyperglycemia. Nat Med 24: 518-524, 2018.
2385 248. Chung JO, Koutsari C, Blachnio-Zabielska AU, Hames KC, and Jensen MD. Intramyocellular
2386 Ceramides: Subcellular Concentrations and Fractional De Novo Synthesis in Postabsorptive Humans.
2387 Diabetes 66: 2082-2091, 2017.
2388 249. Norris GH, and Blesso CN. Dietary and Endogenous Sphingolipid Metabolism in Chronic
2389 Inflammation. Nutrients 9: 2017.
96
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2390 250. Johnson EL, Heaver SL, Waters JL, Kim BI, Bretin A, Goodman AL, Gewirtz AT, Worgall TS, and
2391 Ley RE. Sphingolipids produced by gut bacteria enter host metabolic pathways impacting ceramide
2392 levels. Nat Commun 11: 2471, 2020.
2393 251. Nicholson RJ, Norris MK, Poss AM, Holland WL, and Summers SA. The Lard Works in
2394 Mysterious Ways: Ceramides in Nutrition-Linked Chronic Disease. Annu Rev Nutr 2022.
2395 252. Larsen PJ, and Tennagels N. On ceramides, other sphingolipids and impaired glucose
2396 homeostasis. Mol Metab 3: 252-260, 2014.
2397 253. Jiang C, Xie C, Li F, Zhang L, Nichols RG, Krausz KW, Cai J, Qi Y, Fang ZZ, Takahashi S, Tanaka N,
2398 Desai D, Amin SG, Albert I, Patterson AD, and Gonzalez FJ. Intestinal farnesoid X receptor signaling
2399 promotes nonalcoholic fatty liver disease. The Journal of clinical investigation 2014.
2400 254. Jiang J, Ma Y, Liu Y, Lu D, Gao X, Krausz KW, Desai D, Amin SG, Patterson AD, Gonzalez FJ, and
2401 Xie C. Glycine-beta-muricholic acid antagonizes the intestinal farnesoid X receptor-ceramide axis and
2402 ameliorates NASH in mice. Hepatol Commun 6: 3363-3378, 2022.
2403 255. Wu Q, Sun L, Hu X, Wang X, Xu F, Chen B, Liang X, Xia J, Wang P, Aibara D, Zhang S, Zeng G,
2404 Yun C, Yan Y, Zhu Y, Bustin M, Zhang S, Gonzalez FJ, and Jiang C. Suppressing the intestinal farnesoid X
2405 receptor/sphingomyelin phosphodiesterase 3 axis decreases atherosclerosis. J Clin Invest 131: 2021.
2406 256. Gonzalez FJ, Jiang C, Bisson WH, and Patterson AD. Inhibition of farnesoid X receptor signaling
2407 shows beneficial effects in human obesity. J Hepatol 62: 1234-1236, 2015.
2408 257. Jiang C, Xie C, Li F, Zhang L, Nichols RG, Krausz KW, Cai J, Qi Y, Fang ZZ, Takahashi S, Tanaka N,
2409 Desai D, Amin SG, Albert I, Patterson AD, and Gonzalez FJ. Intestinal farnesoid X receptor signaling
2410 promotes nonalcoholic fatty liver disease. J Clin Invest 125: 386-402, 2015.
2411 258. Jiang C, Xie C, Lv Y, Li J, Krausz KW, Shi J, Brocker CN, Desai D, Amin SG, Bisson WH, Liu Y,
2412 Gavrilova O, Patterson AD, and Gonzalez FJ. Intestine-selective farnesoid X receptor inhibition improves
2413 obesity-related metabolic dysfunction. Nat Commun 6: 10166, 2015.
2414 259. Xie C, Yagai T, Luo Y, Liang X, Chen T, Wang Q, Sun D, Zhao J, Ramakrishnan SK, Sun L, Jiang C,
2415 Xue X, Tian Y, Krausz KW, Patterson AD, Shah YM, Wu Y, Jiang C, and Gonzalez FJ. Activation of
2416 intestinal hypoxia-inducible factor 2alpha during obesity contributes to hepatic steatosis. Nat Med 23:
2417 1298-1308, 2017.
2418 260. Zhang X, Zhang Y, Wang P, Zhang SY, Dong Y, Zeng G, Yan Y, Sun L, Wu Q, Liu H, Liu B, Kong W,
2419 Wang X, and Jiang C. Adipocyte Hypoxia-Inducible Factor 2alpha Suppresses Atherosclerosis by
2420 Promoting Adipose Ceramide Catabolism. Cell Metab 30: 937-951 e935, 2019.
2421 261. Luo Y, Yang S, Wu X, Takahashi S, Sun L, Cai J, Krausz KW, Guo X, Dias HB, Gavrilova O, Xie C,
2422 Jiang C, Liu W, and Gonzalez FJ. Intestinal MYC modulates obesity-related metabolic dysfunction. Nat
2423 Metab 3: 923-939, 2021.
2424 262. de Mello VD, Lankinen M, Schwab U, Kolehmainen M, Lehto S, Seppanen-Laakso T, Oresic M,
2425 Pulkkinen L, Uusitupa M, and Erkkila AT. Link between plasma ceramides, inflammation and insulin
2426 resistance: association with serum IL-6 concentration in patients with coronary heart disease.
2427 Diabetologia 52: 2612-2615, 2009.
2428 263. Majumdar I, and Mastrandrea LD. Serum sphingolipids and inflammatory mediators in
2429 adolescents at risk for metabolic syndrome. Endocrine 41: 442-449, 2012.
2430 264. Hotamisligil GS. Inflammation and metabolic disorders. Nature 444: 860-867, 2006.
2431 265. Schilling JD, Machkovech HM, He L, Sidhu R, Fujiwara H, Weber K, Ory DS, and Schaffer JE.
2432 Palmitate and lipopolysaccharide trigger synergistic ceramide production in primary macrophages. J Biol
2433 Chem 288: 2923-2932, 2013.
2434 266. Shi H, Kokoeva MV, Inouye K, Tzameli I, Yin H, and Flier JS. TLR4 links innate immunity and fatty
2435 acid-induced insulin resistance. J Clin Invest 116: 3015-3025, 2006.
2436 267. Lancaster GI, Langley KG, Berglund NA, Kammoun HL, Reibe S, Estevez E, Weir J, Mellett NA,
2437 Pernes G, Conway JRW, Lee MKS, Timpson P, Murphy AJ, Masters SL, Gerondakis S, Bartonicek N,
97
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2438 Kaczorowski DC, Dinger ME, Meikle PJ, Bond PJ, and Febbraio MA. Evidence that TLR4 Is Not a
2439 Receptor for Saturated Fatty Acids but Mediates Lipid-Induced Inflammation by Reprogramming
2440 Macrophage Metabolism. Cell Metab 27: 1096-1110 e1095, 2018.
2441 268. Sims K, Haynes CA, Kelly S, Allegood JC, Wang E, Momin A, Leipelt M, Reichart D, Glass CK,
2442 Sullards MC, and Merrill AH, Jr. Kdo2-lipid A, a TLR4-specific agonist, induces de novo sphingolipid
2443 biosynthesis in RAW264.7 macrophages, which is essential for induction of autophagy. J Biol Chem 285:
2444 38568-38579, 2010.
2445 269. Vandanmagsar B, Youm YH, Ravussin A, Galgani JE, Stadler K, Mynatt RL, Ravussin E, Stephens
2446 JM, and Dixit VD. The NLRP3 inflammasome instigates obesity-induced inflammation and insulin
2447 resistance. Nat Med 17: 179-188, 2011.
2448 270. Villa NY, Kupchak BR, Garitaonandia I, Smith JL, Alonso E, Alford C, Cowart LA, Hannun YA, and
2449 Lyons TJ. Sphingolipids function as downstream effectors of a fungal PAQR. Mol Pharmacol 75: 866-875,
2450 2009.
2451 271. Kupchak BR, Garitaonandia I, Villa NY, Smith JL, and Lyons TJ. Antagonism of human
2452 adiponectin receptors and their membrane progesterone receptor paralogs by TNFalpha and a
2453 ceramidase inhibitor. Biochemistry 48: 5504-5506, 2009.
2454 272. Scherer LGSaPE. Metabolic Messengers: adiponectin. Nature Metabolism 1: 334-339, 2019.
2455 273. Yamauchi T, Kamon J, Minokoshi Y, Ito Y, Waki H, Uchida S, Yamashita S, Noda M, Kita S, Ueki
2456 K, Eto K, Akanuma Y, Froguel P, Foufelle F, Ferre P, Carling D, Kimura S, Nagai R, Kahn BB, and
2457 Kadowaki T. Adiponectin stimulates glucose utilization and fatty-acid oxidation by activating AMP-
2458 activated protein kinase. Nat Med 8: 1288-1295, 2002.
2459 274. Tanabe H, Fujii Y, Okada-Iwabu M, Iwabu M, Nakamura Y, Hosaka T, Motoyama K, Ikeda M,
2460 Wakiyama M, Terada T, Ohsawa N, Hato M, Ogasawara S, Hino T, Murata T, Iwata S, Hirata K, Kawano
2461 Y, Yamamoto M, Kimura-Someya T, Shirouzu M, Yamauchi T, Kadowaki T, and Yokoyama S. Crystal
2462 structures of the human adiponectin receptors. Nature 520: 312-316, 2015.
2463 275. Holland WL, and Scherer PE. Structural biology: Receptors grease the metabolic wheels. Nature
2464 544: 42-44, 2017.
2465 276. Xia JY, Holland WL, Kusminski CM, Sun K, Sharma AX, Pearson MJ, Sifuentes AJ, McDonald JG,
2466 Gordillo R, and Scherer PE. Targeted Induction of Ceramide Degradation Leads to Improved Systemic
2467 Metabolism and Reduced Hepatic Steatosis. Cell Metab 22: 266-278, 2015.
2468 277. Simeone CA, Wilkerson JL, Poss AM, Banks JA, Varre JV, Guevara JL, Hernandez EJ, Gorsi B,
2469 Atkinson DL, Turapov T, Frodsham SG, Morales JCF, O'Neil K, Moore B, Yandell M, Summers SA,
2470 Krolewski AS, Holland WL, and Pezzolesi MG. A dominant negative ADIPOQ mutation in a diabetic
2471 family with renal disease, hypoadiponectinemia, and hyperceramidemia. NPJ Genom Med 7: 43, 2022.
2472 278. Pei J, Millay DP, Olson EN, and Grishin NV. CREST--a large and diverse superfamily of putative
2473 transmembrane hydrolases. Biol Direct 6: 37, 2011.
2474 279. Pedersen L, Panahandeh P, Siraji MI, Knappskog S, Lonning PE, Gordillo R, Scherer PE, Molven
2475 A, Teigen K, and Halberg N. Golgi-Localized PAQR4 Mediates Antiapoptotic Ceramidase Activity in
2476 Breast Cancer. Cancer Res 80: 2163-2174, 2020.
2477 280. Demirkan A, van Duijn CM, Ugocsai P, Isaacs A, Pramstaller PP, Liebisch G, Wilson JF,
2478 Johansson A, Rudan I, Aulchenko YS, Kirichenko AV, Janssens AC, Jansen RC, Gnewuch C, Domingues
2479 FS, Pattaro C, Wild SH, Jonasson I, Polasek O, Zorkoltseva IV, Hofman A, Karssen LC, Struchalin M,
2480 Floyd J, Igl W, Biloglav Z, Broer L, Pfeufer A, Pichler I, Campbell S, Zaboli G, Kolcic I, Rivadeneira F,
2481 Huffman J, Hastie ND, Uitterlinden A, Franke L, Franklin CS, Vitart V, Consortium D, Nelson CP, Preuss
2482 M, Consortium CA, Bis JC, O'Donnell CJ, Franceschini N, Consortium C, Witteman JC, Axenovich T,
2483 Oostra BA, Meitinger T, Hicks AA, Hayward C, Wright AF, Gyllensten U, Campbell H, Schmitz G, and
2484 consortium E. Genome-wide association study identifies novel loci associated with circulating phospho-
2485 and sphingolipid concentrations. PLoS Genet 8: e1002490, 2012.
98
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2486 281. Linke V, Overmyer KA, Miller IJ, Brademan DR, Hutchins PD, Trujillo EA, Reddy TR, Russell JD,
2487 Cushing EM, Schueler KL, Stapleton DS, Rabaglia ME, Keller MP, Gatti DM, Keele GR, Pham D, Broman
2488 KW, Churchill GA, Attie AD, and Coon JJ. A large-scale genome-lipid association map guides lipid
2489 identification. Nat Metab 2: 1149-1162, 2020.
2490 282. Badshah H, Ali T, and Kim MO. Osmotin attenuates LPS-induced neuroinflammation and
2491 memory impairments via the TLR4/NFkappaB signaling pathway. Sci Rep 6: 24493, 2016.
2492 283. Takahashi Y, Watanabe R, Sato Y, Ozawa N, Kojima M, Watanabe-Kominato K, Shirai R, Sato K,
2493 Hirano T, and Watanabe T. Novel phytopeptide osmotin mimics preventive effects of adiponectin on
2494 vascular inflammation and atherosclerosis. Metabolism 83: 128-138, 2018.
2495 284. Ahmad A, Ali T, Kim MW, Khan A, Jo MH, Rehman SU, Khan MS, Abid NB, Khan M, Ullah R, Jo
2496 MG, and Kim MO. Adiponectin homolog novel osmotin protects obesity/diabetes-induced NAFLD by
2497 upregulating AdipoRs/PPARalpha signaling in ob/ob and db/db transgenic mouse models. Metabolism
2498 90: 31-43, 2019.
2499 285. Summers SA, Chaurasia B, and Holland WL. Metabolic Messengers: ceramides. Nat Metab 1:
2500 1051-1058, 2019.
2501 286. Hammerschmidt P, Ostkotte D, Nolte H, Gerl MJ, Jais A, Brunner HL, Sprenger H-G, Awazawa
2502 M, Nicholls HT, and Turpin-Nolan SM. CerS6-derived sphingolipids interact with Mff and promote
2503 mitochondrial fragmentation in obesity. Cell 177: 1536-1552. e1523, 2019.
2504 287. Chen Q, Denard B, Lee CE, Han S, Ye JS, and Ye J. Inverting the Topology of a Transmembrane
2505 Protein by Regulating the Translocation of the First Transmembrane Helix. Mol Cell 63: 567-578, 2016.
2506 288. Herrmann JL, Bruckheimer E, and McDonnell TJ. Cell death signal transduction and Bcl-2
2507 function. Biochem Soc Trans 24: 1059-1065, 1996.
2508 289. Smyth MJ, Perry DK, Zhang J, Poirier GG, Hannun YA, and Obeid LM. prICE: a downstream
2509 target for ceramide-induced apoptosis and for the inhibitory action of Bcl-2. Biochem J 316 ( Pt 1): 25-28,
2510 1996.
2511 290. Zhang J, Alter N, Reed JC, Borner C, Obeid LM, and Hannun YA. Bcl-2 interrupts the ceramide-
2512 mediated pathway of cell death. Proc Natl Acad Sci U S A 93: 5325-5328, 1996.
2513 291. Wieder T, Geilen CC, Kolter T, Sadeghlar F, Sandhoff K, Brossmer R, Ihrig P, Perry D, Orfanos
2514 CE, and Hannun YA. Bcl-2 antagonizes apoptotic cell death induced by two new ceramide analogues.
2515 FEBS Lett 411: 260-264, 1997.
2516 292. Richter C, and Ghafourifar P. Ceramide induces cytochrome c release from isolated
2517 mitochondria. Biochem Soc Symp 66: 27-31, 1999.
2518 293. Colombini M. Ceramide channels and their role in mitochondria-mediated apoptosis. Biochim
2519 Biophys Acta 1797: 1239-1244, 2010.
2520 294. Ganesan V, Perera MN, Colombini D, Datskovskiy D, Chadha K, and Colombini M. Ceramide
2521 and activated Bax act synergistically to permeabilize the mitochondrial outer membrane. Apoptosis 15:
2522 553-562, 2010.
2523 295. Alonso A, and Goni FM. The Physical Properties of Ceramides in Membranes. Annu Rev Biophys
2524 47: 633-654, 2018.
2525 296. Dingjan T, and Futerman AH. The role of the 'sphingoid motif' in shaping the molecular
2526 interactions of sphingolipids in biomembranes. Biochim Biophys Acta Biomembr 1863: 183701, 2021.
2527 297. Ramstedt B, and Slotte JP. Membrane properties of sphingomyelins. FEBS Lett 531: 33-37, 2002.
2528 298. Simons K, and Ikonen E. Functional rafts in cell membranes. Nature 387: 569-572, 1997.
2529 299. Kolesnick RN, Goni FM, and Alonso A. Compartmentalization of ceramide signaling: physical
2530 foundations and biological effects. J Cell Physiol 184: 285-300, 2000.
2531 300. Brown DA, and London E. Functions of lipid rafts in biological membranes. Annu Rev Cell Dev
2532 Biol 14: 111-136, 1998.
99
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2533 301. Kinoshita M, and Matsumori N. Inimitable Impacts of Ceramides on Lipid Rafts Formed in
2534 Artificial and Natural Cell Membranes. Membranes (Basel) 12: 2022.
2535 302. Xu X, Bittman R, Duportail G, Heissler D, Vilcheze C, and London E. Effect of the structure of
2536 natural sterols and sphingolipids on the formation of ordered sphingolipid/sterol domains (rafts).
2537 Comparison of cholesterol to plant, fungal, and disease-associated sterols and comparison of
2538 sphingomyelin, cerebrosides, and ceramide. J Biol Chem 276: 33540-33546, 2001.
2539 303. Megha, Bakht O, and London E. Cholesterol precursors stabilize ordinary and ceramide-rich
2540 ordered lipid domains (lipid rafts) to different degrees. Implications for the Bloch hypothesis and sterol
2541 biosynthesis disorders. J Biol Chem 281: 21903-21913, 2006.
2542 304. Megha, and London E. Ceramide selectively displaces cholesterol from ordered lipid domains
2543 (rafts): implications for lipid raft structure and function. J Biol Chem 279: 9997-10004, 2004.
2544 305. Holopainen JM, Subramanian M, and Kinnunen PK. Sphingomyelinase induces lipid
2545 microdomain formation in a fluid phosphatidylcholine/sphingomyelin membrane. Biochemistry 37:
2546 17562-17570, 1998.
2547 306. Nurminen TA, Holopainen JM, Zhao H, and Kinnunen PK. Observation of topical catalysis by
2548 sphingomyelinase coupled to microspheres. J Am Chem Soc 124: 12129-12134, 2002.
2549 307. Veiga MP, Arrondo JL, Goni FM, and Alonso A. Ceramides in phospholipid membranes: effects
2550 on bilayer stability and transition to nonlamellar phases. Biophys J 76: 342-350, 1999.
2551 308. Dobrowsky RT, Kamibayashi C, Mumby MC, and Hannun YA. Ceramide activates heterotrimeric
2552 protein phosphatase 2A. J Biol Chem 268: 15523-15530, 1993.
2553 309. Law B, and Rossie S. The dimeric and catalytic subunit forms of protein phosphatase 2A from rat
2554 brain are stimulated by C2-ceramide. J Biol Chem 270: 12808-12813, 1995.
2555 310. Galve-Roperh I, Haro A, and Diaz-Laviada I. Ceramide-induced translocation of protein kinase C
2556 zeta in primary cultures of astrocytes. FEBS Lett 415: 271-274, 1997.
2557 311. Bourbon NA, Yun J, and Kester M. Ceramide directly activates protein kinase C zeta to regulate
2558 a stress-activated protein kinase signaling complex. J Biol Chem 275: 35617-35623, 2000.
2559 312. Mukhopadhyay A, Saddoughi SA, Song P, Sultan I, Ponnusamy S, Senkal CE, Snook CF, Arnold
2560 HK, Sears RC, Hannun YA, and Ogretmen B. Direct interaction between the inhibitor 2 and ceramide via
2561 sphingolipid-protein binding is involved in the regulation of protein phosphatase 2A activity and
2562 signaling. FASEB J 23: 751-763, 2009.
2563 313. Hajduch E, Turban S, Le Liepvre X, Le Lay S, Lipina C, Dimopoulos N, Dugail I, and Hundal HS.
2564 Targeting of PKCzeta and PKB to caveolin-enriched microdomains represents a crucial step underpinning
2565 the disruption in PKB-directed signalling by ceramide. Biochem J 410: 369-379, 2008.
2566 314. Pohl J, Ring A, Ehehalt R, Schulze-Bergkamen H, Schad A, Verkade P, and Stremmel W. Long-
2567 chain fatty acid uptake into adipocytes depends on lipid raft function. Biochemistry 43: 4179-4187, 2004.
2568 315. Pohl J, Ring A, and Stremmel W. Uptake of long-chain fatty acids in HepG2 cells involves
2569 caveolae: analysis of a novel pathway. J Lipid Res 43: 1390-1399, 2002.
2570 316. Meshulam T, Breen MR, Liu L, Parton RG, and Pilch PF. Caveolins/caveolae protect adipocytes
2571 from fatty acid-mediated lipotoxicity. J Lipid Res 52: 1526-1532, 2011.
2572 317. Chaurasia B, Tippetts TS, Mayoral Monibas R, Liu J, Li Y, Wang L, Wilkerson JL, Sweeney CR,
2573 Pereira RF, Sumida DH, Maschek JA, Cox JE, Kaddai V, Lancaster GI, Siddique MM, Poss A, Pearson M,
2574 Satapati S, Zhou H, McLaren DG, Previs SF, Chen Y, Qian Y, Petrov A, Wu M, Shen X, Yao J, Nunes CN,
2575 Howard AD, Wang L, Erion MD, Rutter J, Holland WL, Kelley DE, and Summers SA. Targeting a ceramide
2576 double bond improves insulin resistance and hepatic steatosis. Science 365: 386-392, 2019.
2577 318. Xia JY, Holland WL, Kusminski CM, Sun K, Sharma AX, Pearson MJ, Sifuentes AJ, McDonald JG,
2578 Gordillo R, and Scherer PE. Targeted Induction of Ceramide Degradation Leads to Improved Systemic
2579 Metabolism and Reduced Hepatic Steatosis. Cell Metab 22: 266-278, 2015.
100
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2580 319. Powell DJ, Hajduch E, Kular G, and Hundal HS. Ceramide disables 3-phosphoinositide binding to
2581 the pleckstrin homology domain of protein kinase B (PKB)/Akt by a PKCzeta-dependent mechanism. Mol
2582 Cell Biol 23: 7794-7808, 2003.
2583 320. Bourbon NA, Sandirasegarane L, and Kester M. Ceramide-induced inhibition of Akt is mediated
2584 through protein kinase Czeta: implications for growth arrest. J Biol Chem 277: 3286-3292, 2002.
2585 321. Taniguchi CM, Kondo T, Sajan M, Luo J, Bronson R, Asano T, Farese R, Cantley LC, and Kahn CR.
2586 Divergent regulation of hepatic glucose and lipid metabolism by phosphoinositide 3-kinase via Akt and
2587 PKClambda/zeta. Cell Metab 3: 343-353, 2006.
2588 322. Chaurasia B, Kaddai VA, Lancaster GI, Henstridge DC, Sriram S, Galam DL, Gopalan V, Prakash
2589 KN, Velan SS, Bulchand S, Tsong TJ, Wang M, Siddique MM, Yuguang G, Sigmundsson K, Mellet NA,
2590 Weir JM, Meikle PJ, Bin MYMS, Shabbir A, Shayman JA, Hirabayashi Y, Shiow ST, Sugii S, and Summers
2591 SA. Adipocyte Ceramides Regulate Subcutaneous Adipose Browning, Inflammation, and Metabolism.
2592 Cell Metab 24: 820-834, 2016.
2593 323. Summers SA, Garza LA, Zhou H, and Birnbaum MJ. Regulation of insulin-stimulated glucose
2594 transporter GLUT4 translocation and Akt kinase activity by ceramide. Mol Cell Biol 18: 5457-5464, 1998.
2595 324. Wang CN, O'Brien L, and Brindley DN. Effects of cell-permeable ceramides and tumor necrosis
2596 factor-alpha on insulin signaling and glucose uptake in 3T3-L1 adipocytes. Diabetes 47: 24-31, 1998.
2597 325. Blouin CM, Prado C, Takane KK, Lasnier F, Garcia-Ocana A, Ferre P, Dugail I, and Hajduch E.
2598 Plasma membrane subdomain compartmentalization contributes to distinct mechanisms of ceramide
2599 action on insulin signaling. Diabetes 59: 600-610, 2010.
2600 326. Kim SW, Kim HJ, Chun YJ, and Kim MY. Ceramide produces apoptosis through induction of
2601 p27(kip1) by protein phosphatase 2A-dependent Akt dephosphorylation in PC-3 prostate cancer cells. J
2602 Toxicol Environ Health A 73: 1465-1476, 2010.
2603 327. Yu T, Li J, Qiu Y, and Sun H. 1-phenyl-2-decanoylamino-3-morpholino-1-propanol (PDMP)
2604 facilitates curcumin-induced melanoma cell apoptosis by enhancing ceramide accumulation, JNK
2605 activation, and inhibiting PI3K/AKT activation. Mol Cell Biochem 361: 47-54, 2012.
2606 328. Tagaram HR, Divittore NA, Barth BM, Kaiser JM, Avella D, Kimchi ET, Jiang Y, Isom HC, Kester
2607 M, and Staveley-O'Carroll KF. Nanoliposomal ceramide prevents in vivo growth of hepatocellular
2608 carcinoma. Gut 60: 695-701, 2011.
2609 329. Hresko RC, and Mueckler M. mTOR.RICTOR is the Ser473 kinase for Akt/protein kinase B in 3T3-
2610 L1 adipocytes. J Biol Chem 280: 40406-40416, 2005.
2611 330. Sarbassov DD, Guertin DA, Ali SM, and Sabatini DM. Phosphorylation and regulation of Akt/PKB
2612 by the rictor-mTOR complex. Science 307: 1098-1101, 2005.
2613 331. Stratford S, Hoehn KL, Liu F, and Summers SA. Regulation of insulin action by ceramide: dual
2614 mechanisms linking ceramide accumulation to the inhibition of Akt/protein kinase B. J Biol Chem 279:
2615 36608-36615, 2004.
2616 332. Fox TE, Houck KL, O'Neill SM, Nagarajan M, Stover TC, Pomianowski PT, Unal O, Yun JK, Naides
2617 SJ, and Kester M. Ceramide recruits and activates protein kinase C zeta (PKC zeta) within structured
2618 membrane microdomains. J Biol Chem 282: 12450-12457, 2007.
2619 333. Gudz TI, Tserng KY, and Hoppel CL. Direct inhibition of mitochondrial respiratory chain complex
2620 III by cell-permeable ceramide. J Biol Chem 272: 24154-24158, 1997.
2621 334. Di Paola M, Cocco T, and Lorusso M. Ceramide interaction with the respiratory chain of heart
2622 mitochondria. Biochemistry 39: 6660-6668, 2000.
2623 335. Zigdon H, Kogot-Levin A, Park JW, Goldschmidt R, Kelly S, Merrill AH, Jr., Scherz A, Pewzner-
2624 Jung Y, Saada A, and Futerman AH. Ablation of ceramide synthase 2 causes chronic oxidative stress due
2625 to disruption of the mitochondrial respiratory chain. J Biol Chem 288: 4947-4956, 2013.
101
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2626 336. Park M, Kaddai V, Ching J, Fridianto KT, Sieli RJ, Sugii S, and Summers SA. A Role for Ceramides,
2627 but Not Sphingomyelins, as Antagonists of Insulin Signaling and Mitochondrial Metabolism in C2C12
2628 Myotubes. J Biol Chem 291: 23978-23988, 2016.
2629 337. Diaz-Vegas A, Madsen S, Cooke KC, Carroll L, Khor JXY, Turner N, Lim XY, Astore MA, Morris J,
2630 Don A, Garfield A, Zarini S, Zemski Berry KA, Ryan A, Bergman BC, Brozinick JT, James DE, and
2631 Burchfield JG. Mitochondrial electron transport chain, ceramide and Coenzyme Q are linked in a
2632 pathway that drives insulin resistance in skeletal muscle. bioRxiv 2023.
2633 338. Hammerschmidt P, Ostkotte D, Nolte H, Gerl MJ, Jais A, Brunner HL, Sprenger HG, Awazawa
2634 M, Nicholls HT, Turpin-Nolan SM, Langer T, Kruger M, Brugger B, and Bruning JC. CerS6-Derived
2635 Sphingolipids Interact with Mff and Promote Mitochondrial Fragmentation in Obesity. Cell 177: 1536-
2636 1552 e1523, 2019.
2637 339. Turpin-Nolan SM, Hammerschmidt P, Chen W, Jais A, Timper K, Awazawa M, Brodesser S, and
2638 Bruning JC. CerS1-Derived C18:0 Ceramide in Skeletal Muscle Promotes Obesity-Induced Insulin
2639 Resistance. Cell Rep 26: 1-10 e17, 2019.
2640 340. Turner N, Lim XY, Toop HD, Osborne B, Brandon AE, Taylor EN, Fiveash CE, Govindaraju H, Teo
2641 JD, McEwen HP, Couttas TA, Butler SM, Das A, Kowalski GM, Bruce CR, Hoehn KL, Fath T, Schmitz-
2642 Peiffer C, Cooney GJ, Montgomery MK, Morris JC, and Don AS. A selective inhibitor of ceramide
2643 synthase 1 reveals a novel role in fat metabolism. Nat Commun 9: 3165, 2018.
2644 341. Smith ME, Tippetts TS, Brassfield ES, Tucker BJ, Ockey A, Swensen AC, Anthonymuthu TS,
2645 Washburn TD, Kane DA, Prince JT, and Bikman BT. Mitochondrial fission mediates ceramide-induced
2646 metabolic disruption in skeletal muscle. Biochem J 456: 427-439, 2013.
2647 342. Kota V, Szulc ZM, and Hama H. Identification of C(6) -ceramide-interacting proteins in D6P2T
2648 Schwannoma cells. Proteomics 12: 2179-2184, 2012.
2649 343. Kogot-Levin A, and Saada A. Ceramide and the mitochondrial respiratory chain. Biochimie 100:
2650 88-94, 2014.
2651 344. Abdel Shakor AB, Kwiatkowska K, and Sobota A. Cell surface ceramide generation precedes and
2652 controls FcgammaRII clustering and phosphorylation in rafts. J Biol Chem 279: 36778-36787, 2004.
2653 345. Dumitru CA, and Gulbins E. TRAIL activates acid sphingomyelinase via a redox mechanism and
2654 releases ceramide to trigger apoptosis. Oncogene 25: 5612-5625, 2006.
2655 346. Goggel R, Winoto-Morbach S, Vielhaber G, Imai Y, Lindner K, Brade L, Brade H, Ehlers S,
2656 Slutsky AS, Schutze S, Gulbins E, and Uhlig S. PAF-mediated pulmonary edema: a new role for acid
2657 sphingomyelinase and ceramide. Nat Med 10: 155-160, 2004.
2658 347. Grassme H, Jekle A, Riehle A, Schwarz H, Berger J, Sandhoff K, Kolesnick R, and Gulbins E. CD95
2659 signaling via ceramide-rich membrane rafts. J Biol Chem 276: 20589-20596, 2001.
2660 348. Grassme H, Jendrossek V, Bock J, Riehle A, and Gulbins E. Ceramide-rich membrane rafts
2661 mediate CD40 clustering. J Immunol 168: 298-307, 2002.
2662 349. Grassme H, Schwarz H, and Gulbins E. Molecular mechanisms of ceramide-mediated CD95
2663 clustering. Biochem Biophys Res Commun 284: 1016-1030, 2001.
2664 350. Pfeiffer A, Bottcher A, Orso E, Kapinsky M, Nagy P, Bodnar A, Spreitzer I, Liebisch G, Drobnik
2665 W, Gempel K, Horn M, Holmer S, Hartung T, Multhoff G, Schutz G, Schindler H, Ulmer AJ, Heine H,
2666 Stelter F, Schutt C, Rothe G, Szollosi J, Damjanovich S, and Schmitz G. Lipopolysaccharide and ceramide
2667 docking to CD14 provokes ligand-specific receptor clustering in rafts. Eur J Immunol 31: 3153-3164,
2668 2001.
2669 351. Grassme H, Bock J, Kun J, and Gulbins E. Clustering of CD40 ligand is required to form a
2670 functional contact with CD40. J Biol Chem 277: 30289-30299, 2002.
2671 352. Garcia-Barros M, Paris F, Cordon-Cardo C, Lyden D, Rafii S, Haimovitz-Friedman A, Fuks Z, and
2672 Kolesnick R. Tumor response to radiotherapy regulated by endothelial cell apoptosis. Science 300: 1155-
2673 1159, 2003.
102
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2674 353. Charruyer A, Grazide S, Bezombes C, Muller S, Laurent G, and Jaffrezou JP. UV-C light induces
2675 raft-associated acid sphingomyelinase and JNK activation and translocation independently on a nuclear
2676 signal. J Biol Chem 280: 19196-19204, 2005.
2677 354. Grassme H, Riehle A, Wilker B, and Gulbins E. Rhinoviruses infect human epithelial cells via
2678 ceramide-enriched membrane platforms. J Biol Chem 280: 26256-26262, 2005.
2679 355. Kashkar H, Wiegmann K, Yazdanpanah B, Haubert D, and Kronke M. Acid sphingomyelinase is
2680 indispensable for UV light-induced Bax conformational change at the mitochondrial membrane. J Biol
2681 Chem 280: 20804-20813, 2005.
2682 356. Lacour S, Hammann A, Grazide S, Lagadic-Gossmann D, Athias A, Sergent O, Laurent G,
2683 Gambert P, Solary E, and Dimanche-Boitrel MT. Cisplatin-induced CD95 redistribution into membrane
2684 lipid rafts of HT29 human colon cancer cells. Cancer Res 64: 3593-3598, 2004.
2685 357. Paris F, Fuks Z, Kang A, Capodieci P, Juan G, Ehleiter D, Haimovitz-Friedman A, Cordon-Cardo
2686 C, and Kolesnick R. Endothelial apoptosis as the primary lesion initiating intestinal radiation damage in
2687 mice. Science 293: 293-297, 2001.
2688 358. Rotolo JA, Zhang J, Donepudi M, Lee H, Fuks Z, and Kolesnick R. Caspase-dependent and -
2689 independent activation of acid sphingomyelinase signaling. J Biol Chem 280: 26425-26434, 2005.
2690 359. Zhang Y, Mattjus P, Schmid PC, Dong Z, Zhong S, Ma WY, Brown RE, Bode AM, Schmid HH, and
2691 Dong Z. Involvement of the acid sphingomyelinase pathway in uva-induced apoptosis. J Biol Chem 276:
2692 11775-11782, 2001.
2693 360. Morad SA, and Cabot MC. Ceramide-orchestrated signalling in cancer cells. Nat Rev Cancer 13:
2694 51-65, 2013.
2695 361. Voelkel-Johnson C, Hannun YA, and El-Zawahry A. Resistance to TRAIL is associated with
2696 defects in ceramide signaling that can be overcome by exogenous C6-ceramide without requiring down-
2697 regulation of cellular FLICE inhibitory protein. Mol Cancer Ther 4: 1320-1327, 2005.
2698 362. White-Gilbertson S, Mullen T, Senkal C, Lu P, Ogretmen B, Obeid L, and Voelkel-Johnson C.
2699 Ceramide synthase 6 modulates TRAIL sensitivity and nuclear translocation of active caspase-3 in colon
2700 cancer cells. Oncogene 28: 1132-1141, 2009.
2701 363. Stiban J, Caputo L, and Colombini M. Ceramide synthesis in the endoplasmic reticulum can
2702 permeabilize mitochondria to proapoptotic proteins. J Lipid Res 49: 625-634, 2008.
2703 364. Novgorodov SA, Wu BX, Gudz TI, Bielawski J, Ovchinnikova TV, Hannun YA, and Obeid LM.
2704 Novel pathway of ceramide production in mitochondria: thioesterase and neutral ceramidase produce
2705 ceramide from sphingosine and acyl-CoA. J Biol Chem 286: 25352-25362, 2011.
2706 365. Lee H, Rotolo JA, Mesicek J, Penate-Medina T, Rimner A, Liao WC, Yin X, Ragupathi G, Ehleiter
2707 D, Gulbins E, Zhai D, Reed JC, Haimovitz-Friedman A, Fuks Z, and Kolesnick R. Mitochondrial ceramide-
2708 rich macrodomains functionalize Bax upon irradiation. PLoS One 6: e19783, 2011.
2709 366. Birbes H, El Bawab S, Hannun YA, and Obeid LM. Selective hydrolysis of a mitochondrial pool of
2710 sphingomyelin induces apoptosis. FASEB J 15: 2669-2679, 2001.
2711 367. Jain A, Beutel O, Ebell K, Korneev S, and Holthuis JC. Diverting CERT-mediated ceramide
2712 transport to mitochondria triggers Bax-dependent apoptosis. J Cell Sci 130: 360-371, 2017.
2713 368. Hernandez-Corbacho MJ, Salama MF, Canals D, Senkal CE, and Obeid LM. Sphingolipids in
2714 mitochondria. Biochim Biophys Acta Mol Cell Biol Lipids 1862: 56-68, 2017.
2715 369. Dadsena S, Bockelmann S, Mina JGM, Hassan DG, Korneev S, Razzera G, Jahn H, Niekamp P,
2716 Muller D, Schneider M, Tafesse FG, Marrink SJ, Melo MN, and Holthuis JCM. Ceramides bind VDAC2 to
2717 trigger mitochondrial apoptosis. Nat Commun 10: 1832, 2019.
2718 370. Siskind LJ, Kolesnick RN, and Colombini M. Ceramide forms channels in mitochondrial outer
2719 membranes at physiologically relevant concentrations. Mitochondrion 6: 118-125, 2006.
2720 371. Siskind LJ, Davoody A, Lewin N, Marshall S, and Colombini M. Enlargement and contracture of
2721 C2-ceramide channels. Biophys J 85: 1560-1575, 2003.
103
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2722 372. Siskind LJ, Kolesnick RN, and Colombini M. Ceramide channels increase the permeability of the
2723 mitochondrial outer membrane to small proteins. J Biol Chem 277: 26796-26803, 2002.
2724 373. Siskind LJ, and Colombini M. The lipids C2- and C16-ceramide form large stable channels.
2725 Implications for apoptosis. J Biol Chem 275: 38640-38644, 2000.
2726 374. Stiban J, Fistere D, and Colombini M. Dihydroceramide hinders ceramide channel formation:
2727 Implications on apoptosis. Apoptosis 11: 773-780, 2006.
2728 375. Perera MN, Lin SH, Peterson YK, Bielawska A, Szulc ZM, Bittman R, and Colombini M. Bax and
2729 Bcl-xL exert their regulation on different sites of the ceramide channel. Biochem J 445: 81-91, 2012.
2730 376. Ganesan V, and Colombini M. Regulation of ceramide channels by Bcl-2 family proteins. FEBS
2731 Lett 584: 2128-2134, 2010.
2732 377. Siskind LJ, Feinstein L, Yu T, Davis JS, Jones D, Choi J, Zuckerman JE, Tan W, Hill RB, Hardwick
2733 JM, and Colombini M. Anti-apoptotic Bcl-2 Family Proteins Disassemble Ceramide Channels. J Biol Chem
2734 283: 6622-6630, 2008.
2735 378. Zhou H, Summers SA, Birnbaum MJ, and Pittman RN. Inhibition of Akt kinase by cell-permeable
2736 ceramide and its implications for ceramide-induced apoptosis. J Biol Chem 273: 16568-16575, 1998.
2737 379. Li Z, Kabir I, Tietelman G, Huan C, Fan J, Worgall T, and Jiang XC. Sphingolipid de novo
2738 biosynthesis is essential for intestine cell survival and barrier function. Cell Death Dis 9: 173, 2018.
2739 380. Ohta E, Ohira T, Matsue K, Ikeda Y, Fujii K, Ohwaki K, Osuka S, Hirabayashi Y, and Sasaki M.
2740 Analysis of development of lesions in mice with serine palmitoyltransferase (SPT) deficiency -Sptlc2
2741 conditional knockout mice. Exp Anim 58: 515-524, 2009.
2742 381. Li Y, Chaurasia B, Rahman MM, Kaddai V, Maschek JA, Berg JA, Wilkerson JL, Mahmassani ZS,
2743 Cox J, Wei P, Meikle PJ, Atkinson D, Wang L, Poss AM, Playdon MC, Tippetts TS, Mousa EM,
2744 Nittayaboon K, Anandh Babu PV, Drummond MJ, Clevers H, Shayman JA, Hirabayashi Y, Holland WL,
2745 Rutter J, Edgar BA, and Summers SA. Ceramides Increase Fatty Acid Utilization in Intestinal Progenitors
2746 to Enhance Stemness and Increase Tumor Risk. Gastroenterology 165: 1136-1150, 2023.
2747 382. Genin MJ, Gonzalez Valcarcel IC, Holloway WG, Lamar J, Mosior M, Hawkins E, Estridge T,
2748 Weidner J, Seng T, Yurek D, Adams LA, Weller J, Reynolds VL, and Brozinick JT. Imidazopyridine and
2749 Pyrazolopiperidine Derivatives as Novel Inhibitors of Serine Palmitoyl Transferase. J Med Chem 59: 5904-
2750 5910, 2016.
2751 383. Beyaz S, Mana MD, Roper J, Kedrin D, Saadatpour A, Hong SJ, Bauer-Rowe KE, Xifaras ME,
2752 Akkad A, Arias E, Pinello L, Katz Y, Shinagare S, Abu-Remaileh M, Mihaylova MM, Lamming DW,
2753 Dogum R, Guo G, Bell GW, Selig M, Nielsen GP, Gupta N, Ferrone CR, Deshpande V, Yuan GC, Orkin SH,
2754 Sabatini DM, and Yilmaz OH. High-fat diet enhances stemness and tumorigenicity of intestinal
2755 progenitors. Nature 531: 53-58, 2016.
2756 384. Mihaylova MM, Cheng CW, Cao AQ, Tripathi S, Mana MD, Bauer-Rowe KE, Abu-Remaileh M,
2757 Clavain L, Erdemir A, Lewis CA, Freinkman E, Dickey AS, La Spada AR, Huang Y, Bell GW, Deshpande V,
2758 Carmeliet P, Katajisto P, Sabatini DM, and Yilmaz OH. Fasting Activates Fatty Acid Oxidation to Enhance
2759 Intestinal Stem Cell Function during Homeostasis and Aging. Cell Stem Cell 22: 769-778 e764, 2018.
2760 385. Bensard CL, Wisidagama DR, Olson KA, Berg JA, Krah NM, Schell JC, Nowinski SM, Fogarty S,
2761 Bott AJ, Wei P, Dove KK, Tanner JM, Panic V, Cluntun A, Lettlova S, Earl CS, Namnath DF, Vazquez-
2762 Arreguin K, Villanueva CJ, Tantin D, Murtaugh LC, Evason KJ, Ducker GS, Thummel CS, and Rutter J.
2763 Regulation of Tumor Initiation by the Mitochondrial Pyruvate Carrier. Cell Metab 31: 284-300 e287,
2764 2020.
2765 386. McGarry JD. What if Minkowski had been ageusic? An alternative angle on diabetes. Science
2766 258: 766-770, 1992.
2767 387. Felig P, Wahren J, Sherwin R, and Palaiologos G. Amino acid and protein metabolism in
2768 diabetes mellitus. Arch Intern Med 137: 507-513, 1977.
104
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2769 388. Huynh K, Barlow CK, Jayawardana KS, Weir JM, Mellett NA, Cinel M, Magliano DJ, Shaw JE,
2770 Drew BG, and Meikle PJ. High-Throughput Plasma Lipidomics: Detailed Mapping of the Associations
2771 with Cardiometabolic Risk Factors. Cell Chem Biol 26: 71-84 e74, 2019.
2772 389. Jensen PN, Fretts AM, Yu C, Hoofnagle AN, Umans JG, Howard BV, Sitlani CM, Siscovick DS,
2773 King IB, Sotoodehnia N, McKnight B, and Lemaitre RN. Circulating sphingolipids, fasting glucose, and
2774 impaired fasting glucose: The Strong Heart Family Study. EBioMedicine 2018.
2775 390. Lemaitre RN, Yu C, Hoofnagle A, Hari N, Jensen PN, Fretts AM, Umans JG, Howard BV, Sitlani
2776 CM, Siscovick DS, King IB, Sotoodehnia N, and McKnight B. Circulating Sphingolipids, Insulin, HOMA-IR,
2777 and HOMA-B: The Strong Heart Family Study. Diabetes 67: 1663-1672, 2018.
2778 391. Chew WS, Torta F, Ji S, Choi H, Begum H, Sim X, Khoo CM, Khoo EYH, Ong WY, Van Dam RM,
2779 Wenk MR, Tai ES, and Herr DR. Large-scale lipidomics identifies associations between plasma
2780 sphingolipids and T2DM incidence. JCI Insight 5: 2019.
2781 392. Wigger L, Cruciani-Guglielmacci C, Nicolas A, Denom J, Fernandez N, Fumeron F, Marques-
2782 Vidal P, Ktorza A, Kramer W, Schulte A, Le Stunff H, Liechti R, Xenarios I, Vollenweider P, Waeber G,
2783 Uphues I, Roussel R, Magnan C, Ibberson M, and Thorens B. Plasma Dihydroceramides Are Diabetes
2784 Susceptibility Biomarker Candidates in Mice and Humans. Cell Rep 18: 2269-2279, 2017.
2785 393. Zarini S, Brozinick JT, Zemski Berry KA, Garfield A, Perreault L, Kerege A, Bui HH, Sanders P,
2786 Siddall P, Kuo MS, and Bergman BC. Serum dihydroceramides correlate with insulin sensitivity in
2787 humans and decrease insulin sensitivity in vitro. J Lipid Res 63: 100270, 2022.
2788 394. Kolak M, Westerbacka J, Velagapudi VR, Wagsater D, Yetukuri L, Makkonen J, Rissanen A,
2789 Hakkinen AM, Lindell M, Bergholm R, Hamsten A, Eriksson P, Fisher RM, Oresic M, and Yki-Jarvinen H.
2790 Adipose tissue inflammation and increased ceramide content characterize subjects with high liver fat
2791 content independent of obesity. Diabetes 56: 1960-1968, 2007.
2792 395. Tonks KT, Coster AC, Christopher MJ, Chaudhuri R, Xu A, Gagnon-Bartsch J, Chisholm DJ, James
2793 DE, Meikle PJ, Greenfield JR, and Samocha-Bonet D. Skeletal muscle and plasma lipidomic signatures of
2794 insulin resistance and overweight/obesity in humans. Obesity (Silver Spring) 24: 908-916, 2016.
2795 396. Bergman BC, Brozinick JT, Strauss A, Bacon S, Kerege A, Bui HH, Sanders P, Siddall P, Wei T,
2796 Thomas MK, Kuo MS, and Perreault L. Muscle sphingolipids during rest and exercise: a C18:0 signature
2797 for insulin resistance in humans. Diabetologia 59: 785-798, 2016.
2798 397. Coen PM, Dube JJ, Amati F, Stefanovic-Racic M, Ferrell RE, Toledo FG, and Goodpaster BH.
2799 Insulin resistance is associated with higher intramyocellular triglycerides in type I but not type II
2800 myocytes concomitant with higher ceramide content. Diabetes 59: 80-88, 2010.
2801 398. Amati F, Dube JJ, Carnero EA, Edreira MM, Chomentowski P, Coen PM, Switzer GE, Bickel PE,
2802 Stefanovic-Racic M, Toledo FG, and Goodpaster BH. Skeletal-Muscle Triglycerides, Diacylglycerols, and
2803 Ceramides in Insulin Resistance: Another Paradox in Endurance-Trained Athletes? Diabetes 2011.
2804 399. Dube JJ, Amati F, Toledo FG, Stefanovic-Racic M, Rossi A, Coen P, and Goodpaster BH. Effects
2805 of weight loss and exercise on insulin resistance, and intramyocellular triacylglycerol, diacylglycerol and
2806 ceramide. Diabetologia 54: 1147-1156, 2011.
2807 400. Coen PM, Hames KC, Leachman EM, DeLany JP, Ritov VB, Menshikova EV, Dube JJ, Stefanovic-
2808 Racic M, Toledo FG, and Goodpaster BH. Reduced skeletal muscle oxidative capacity and elevated
2809 ceramide but not diacylglycerol content in severe obesity. Obesity (Silver Spring) 21: 2362-2371, 2013.
2810 401. Summers SA, and Goodpaster BH. CrossTalk proposal: Intramyocellular ceramide accumulation
2811 does modulate insulin resistance. J Physiol 594: 3167-3170, 2016.
2812 402. Zick Y. Insulin resistance: a phosphorylation-based uncoupling of insulin signaling. Trends Cell
2813 Biol 11: 437-441, 2001.
2814 403. Shimabukuro M, Zhou YT, Levi M, and Unger RH. Fatty acid-induced beta cell apoptosis: a link
2815 between obesity and diabetes. Proc Natl Acad Sci U S A 95: 2498-2502, 1998.
105
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2816 404. Shimabukuro M, Higa M, Zhou YT, Wang MY, Newgard CB, and Unger RH. Lipoapoptosis in
2817 beta-cells of obese prediabetic fa/fa rats. Role of serine palmitoyltransferase overexpression. J Biol
2818 Chem 273: 32487-32490, 1998.
2819 405. Kelpe CL, Moore PC, Parazzoli SD, Wicksteed B, Rhodes CJ, and Poitout V. Palmitate inhibition
2820 of insulin gene expression is mediated at the transcriptional level via ceramide synthesis. J Biol Chem
2821 278: 30015-30021, 2003.
2822 406. Raichur S, Wang ST, Chan PW, Li Y, Ching J, Chaurasia B, Dogra S, Ohman MK, Takeda K, Sugii
2823 S, Pewzner-Jung Y, Futerman AH, and Summers SA. CerS2 Haploinsufficiency Inhibits beta-Oxidation
2824 and Confers Susceptibility to Diet-Induced Steatohepatitis and Insulin Resistance. Cell Metab 20: 919,
2825 2014.
2826 407. Dekker MJ, Baker C, Naples M, Samsoondar J, Zhang R, Qiu W, Sacco J, and Adeli K. Inhibition
2827 of sphingolipid synthesis improves dyslipidemia in the diet-induced hamster model of insulin resistance:
2828 evidence for the role of sphingosine and sphinganine in hepatic VLDL-apoB100 overproduction.
2829 Atherosclerosis 228: 98-109, 2013.
2830 408. Campana M, Bellini L, Rouch C, Rachdi L, Coant N, Butin N, Bandet CL, Philippe E, Meneyrol K,
2831 Kassis N, Dairou J, Hajduch E, Colsch B, Magnan C, and Le Stunff H. Inhibition of central de novo
2832 ceramide synthesis restores insulin signaling in hypothalamus and enhances beta-cell function of obese
2833 Zucker rats. Mol Metab 8: 23-36, 2018.
2834 409. Yang Q, Graham TE, Mody N, Preitner F, Peroni OD, Zabolotny JM, Kotani K, Quadro L, and
2835 Kahn BB. Serum retinol binding protein 4 contributes to insulin resistance in obesity and type 2
2836 diabetes. Nature 436: 356-362, 2005.
2837 410. Preitner F, Mody N, Graham TE, Peroni OD, and Kahn BB. Long-term Fenretinide treatment
2838 prevents high-fat diet-induced obesity, insulin resistance, and hepatic steatosis. Am J Physiol Endocrinol
2839 Metab 297: E1420-1429, 2009.
2840 411. Zheng W, Kollmeyer J, Symolon H, Momin A, Munter E, Wang E, Kelly S, Allegood JC, Liu Y,
2841 Peng Q, Ramaraju H, Sullards MC, Cabot M, and Merrill AH, Jr. Ceramides and other bioactive
2842 sphingolipid backbones in health and disease: lipidomic analysis, metabolism and roles in membrane
2843 structure, dynamics, signaling and autophagy. Biochim Biophys Acta 1758: 1864-1884, 2006.
2844 412. Wang H, Maurer BJ, Liu YY, Wang E, Allegood JC, Kelly S, Symolon H, Liu Y, Merrill AH, Jr.,
2845 Gouaze-Andersson V, Yu JY, Giuliano AE, and Cabot MC. N-(4-Hydroxyphenyl)retinamide increases
2846 dihydroceramide and synergizes with dimethylsphingosine to enhance cancer cell killing. Mol Cancer
2847 Ther 7: 2967-2976, 2008.
2848 413. Blitzer JT, Wang L, and Summers SA. DES1: A Key Driver of Lipotoxicity in Metabolic Disease.
2849 DNA Cell Biol 39: 733-737, 2020.
2850 414. Kummar S, Gutierrez ME, Maurer BJ, Reynolds CP, Kang M, Singh H, Crandon S, Murgo AJ, and
2851 Doroshow JH. Phase I trial of fenretinide lym-x-sorb oral powder in adults with solid tumors and
2852 lymphomas. Anticancer Res 31: 961-966, 2011.
2853 415. Cowan AJ, Stevenson PA, Gooley TA, Frayo SL, Oliveira GR, Smith SD, Green DJ, Roden JE,
2854 Pagel JM, Wood BL, Press OW, and Gopal AK. Results of a phase I-II study of fenretinide and rituximab
2855 for patients with indolent B-cell lymphoma and mantle cell lymphoma. Br J Haematol 176: 583-590,
2856 2017.
2857 416. Pienta KJ, Esper PS, Zwas F, Krzeminski R, and Flaherty LE. Phase II chemoprevention trial of
2858 oral fenretinide in patients at risk for adenocarcinoma of the prostate. Am J Clin Oncol 20: 36-39, 1997.
2859 417. Decensi A, Torrisi R, Gozza A, Severi G, Bertelli G, Fontana V, Pensa F, Carozzo L, Traverso A,
2860 Milone S, Dini D, and Costa A. Effect of fenretinide on bone mineral density and metabolism in women
2861 with early breast cancer. Breast Cancer Res Treat 53: 145-151, 1999.
106
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2862 418. Moore MM, Stockler M, Lim R, Mok TS, Millward M, and Boyer MJ. A phase II study of
2863 fenretinide in patients with hormone refractory prostate cancer: a trial of the Cancer Therapeutics
2864 Research Group. Cancer Chemother Pharmacol 66: 845-850, 2010.
2865 419. Colombo N, Formelli F, Cantu MG, Parma G, Gasco M, Argusti A, Santinelli A, Montironi R,
2866 Cavadini E, Baglietto L, Guerrieri-Gonzaga A, Viale G, and Decensi A. A phase I-II preoperative
2867 biomarker trial of fenretinide in ascitic ovarian cancer. Cancer Epidemiol Biomarkers Prev 15: 1914-1919,
2868 2006.
2869 420. Decensi A, Johansson H, Miceli R, Mariani L, Camerini T, Cavadini E, Di Mauro MG, Barreca A,
2870 Gonzaga AG, Diani S, Sandri MT, De Palo G, and Formelli F. Long-term effects of fenretinide, a retinoic
2871 acid derivative, on the insulin-like growth factor system in women with early breast cancer. Cancer
2872 Epidemiol Biomarkers Prev 10: 1047-1053, 2001.
2873 421. Decensi A, Torrisi R, Bruno S, Costantini M, Curotto A, Nicolo G, Malcangi B, Baglietto L,
2874 Bruttini GP, Gatteschi B, Rondanina G, Varaldo M, Perloff M, Malone WF, and Bruzzi P. Randomized
2875 trial of fenretinide in superficial bladder cancer using DNA flow cytometry as an intermediate end point.
2876 Cancer Epidemiol Biomarkers Prev 9: 1071-1078, 2000.
2877 422. Formelli F, Camerini T, Cavadini E, Appierto V, Villani MG, Costa A, De Palo G, Di Mauro MG,
2878 and Veronesi U. Fenretinide breast cancer prevention trial: drug and retinol plasma levels in relation to
2879 age and disease outcome. Cancer Epidemiol Biomarkers Prev 12: 34-41, 2003.
2880 423. Formelli F, Cavadini E, Luksch R, Garaventa A, Appierto V, and Persiani S. Relationship among
2881 pharmacokinetics and pharmacodynamics of fenretinide and plasma retinol reduction in neuroblastoma
2882 patients. Cancer Chemother Pharmacol 66: 993-998, 2010.
2883 424. Rao RD, Cobleigh MA, Gray R, Graham ML, 2nd, Norton L, Martino S, Budd GT, Ingle JN, and
2884 Wood WC. Phase III double-blind, placebo-controlled, prospective randomized trial of adjuvant
2885 tamoxifen vs. tamoxifen and fenretinide in postmenopausal women with positive receptors (EB193): an
2886 intergroup trial coordinated by the Eastern Cooperative Oncology Group. Med Oncol 28 Suppl 1: S39-47,
2887 2011.
2888 425. Gosejacob D, Jager PS, Vom Dorp K, Frejno M, Carstensen AC, Kohnke M, Degen J, Dormann P,
2889 and Hoch M. Ceramide Synthase 5 Is Essential to Maintain C16:0-Ceramide Pools and Contributes to the
2890 Development of Diet-induced Obesity. J Biol Chem 291: 6989-7003, 2016.
2891 426. Turpin-Nolan SM, Hammerschmidt P, Chen W, Jais A, Timper K, Awazawa M, Brodesser S, and
2892 Bruning JC. CerS1-Derived C(18:0) Ceramide in Skeletal Muscle Promotes Obesity-Induced Insulin
2893 Resistance. Cell Rep 26: 1-10 e17, 2019.
2894 427. Li Z, Zhang H, Liu J, Liang CP, Li Y, Li Y, Teitelman G, Beyer T, Bui HH, Peake DA, Zhang Y,
2895 Sanders PE, Kuo MS, Park TS, Cao G, and Jiang XC. Reducing plasma membrane sphingomyelin increases
2896 insulin sensitivity. Mol Cell Biol 31: 4205-4218, 2011.
2897 428. Bijl N, Sokolovic M, Vrins C, Langeveld M, Moerland PD, Ottenhoff R, van Roomen CP,
2898 Claessen N, Boot RG, Aten J, Groen AK, Aerts JM, and van Eijk M. Modulation of glycosphingolipid
2899 metabolism significantly improves hepatic insulin sensitivity and reverses hepatic steatosis in mice.
2900 Hepatology 50: 1431-1441, 2009.
2901 429. van Eijk M, Aten J, Bijl N, Ottenhoff R, van Roomen CP, Dubbelhuis PF, Seeman I, Ghauharali-
2902 van der Vlugt K, Overkleeft HS, Arbeeny C, Groen AK, and Aerts JM. Reducing glycosphingolipid content
2903 in adipose tissue of obese mice restores insulin sensitivity, adipogenesis and reduces inflammation. PLoS
2904 One 4: e4723, 2009.
2905 430. Zhao H, Przybylska M, Wu IH, Zhang J, Maniatis P, Pacheco J, Piepenhagen P, Copeland D,
2906 Arbeeny C, Shayman JA, Aerts JM, Jiang C, Cheng SH, and Yew NS. Inhibiting glycosphingolipid synthesis
2907 ameliorates hepatic steatosis in obese mice. Hepatology 50: 85-93, 2009.
107
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2908 431. Zhao H, Przybylska M, Wu IH, Zhang J, Siegel C, Komarnitsky S, Yew NS, and Cheng SH.
2909 Inhibiting glycosphingolipid synthesis improves glycemic control and insulin sensitivity in animal models
2910 of type 2 diabetes. Diabetes 56: 1210-1218, 2007.
2911 432. Yamashita T, Hashiramoto A, Haluzik M, Mizukami H, Beck S, Norton A, Kono M, Tsuji S,
2912 Daniotti JL, Werth N, Sandhoff R, Sandhoff K, and Proia RL. Enhanced insulin sensitivity in mice lacking
2913 ganglioside GM3. Proc Natl Acad Sci U S A 100: 3445-3449, 2003.
2914 433. Kabayama K, Sato T, Saito K, Loberto N, Prinetti A, Sonnino S, Kinjo M, Igarashi Y, and Inokuchi
2915 J. Dissociation of the insulin receptor and caveolin-1 complex by ganglioside GM3 in the state of insulin
2916 resistance. Proc Natl Acad Sci U S A 104: 13678-13683, 2007.
2917 434. Chavez JA, Siddique MM, Wang ST, Ching J, Shayman JA, and Summers SA. Ceramides and
2918 glucosylceramides are independent antagonists of insulin signaling. J Biol Chem 289: 723-734, 2014.
2919 435. Chavez JA, Holland WL, Bar J, Sandhoff K, and Summers SA. Acid ceramidase overexpression
2920 prevents the inhibitory effects of saturated fatty acids on insulin signaling. J Biol Chem 280: 20148-
2921 20153, 2005.
2922 436. Chavez JA, Knotts TA, Wang LP, Li G, Dobrowsky RT, Florant GL, and Summers SA. A role for
2923 ceramide, but not diacylglycerol, in the antagonism of insulin signal transduction by saturated fatty
2924 acids. J Biol Chem 278: 10297-10303, 2003.
2925 437. Powell DJ, Turban S, Gray A, Hajduch E, and Hundal HS. Intracellular ceramide synthesis and
2926 protein kinase Czeta activation play an essential role in palmitate-induced insulin resistance in rat L6
2927 skeletal muscle cells. Biochem J 382: 619-629, 2004.
2928 438. Watson ML, Coghlan M, and Hundal HS. Modulating serine palmitoyl transferase (SPT)
2929 expression and activity unveils a crucial role in lipid-induced insulin resistance in rat skeletal muscle
2930 cells. Biochem J 417: 791-801, 2009.
2931 439. Boon J, Hoy AJ, Stark R, Brown RD, Meex RC, Henstridge DC, Schenk S, Meikle PJ, Horowitz JF,
2932 Kingwell BA, Bruce CR, and Watt MJ. Ceramides contained in LDL are elevated in type 2 diabetes and
2933 promote inflammation and skeletal muscle insulin resistance. Diabetes 62: 401-410, 2013.
2934 440. Kohn AD, Summers SA, Birnbaum MJ, and Roth RA. Expression of a constitutively active Akt
2935 Ser/Thr kinase in 3T3-L1 adipocytes stimulates glucose uptake and glucose transporter 4 translocation. J
2936 Biol Chem 271: 31372-31378, 1996.
2937 441. Stratford S, Hoehn KL, and Summers SA. Regulation of insulin signaling by ceramide: two
2938 independent mechanisms linking ceramide accumulation to the inhibition of Akt/Protein Kinase B.
2939 Manuscript Submitted 2003.
2940 442. Stratford S, DeWald DB, and Summers SA. Ceramide dissociates 3'-phosphoinositide production
2941 from pleckstrin homology domain translocation. Biochem J 354: 359-368, 2001.
2942 443. Schubert KM, Scheid MP, and Duronio V. Ceramide inhibits protein kinase B/Akt by promoting
2943 dephosphorylation of serine 473 [In Process Citation]. J Biol Chem 275: 13330-13335, 2000.
2944 444. Zinda MJ, Vlahos CJ, and Lai MT. Ceramide induces the dephosphorylation and inhibition of
2945 constitutively activated Akt in PTEN negative U87mg cells. Biochem Biophys Res Commun 280: 1107-
2946 1115, 2001.
2947 445. Teruel T, Hernandez R, and Lorenzo M. Ceramide mediates insulin resistance by tumor necrosis
2948 factor-alpha in brown adipocytes by maintaining Akt in an inactive dephosphorylated state. Diabetes 50:
2949 2563-2571, 2001.
2950 446. Chavez JA, Knotts TA, Wang LP, Li G, Dobrowsky RT, Florant GL, and Summers SA. A role for
2951 ceramide, but not diacylglycerol, in the antagonism of insulin signal transduction by saturated fatty
2952 acids. J Biol Chem 13: 13, 2003.
2953 447. Salinas M, Lopez-Valdaliso R, Martin D, Alvarez A, and Cuadrado A. Inhibition of PKB/Akt1 by
2954 C2-ceramide involves activation of ceramide- activated protein phosphatase in PC12 cells. Mol Cell
2955 Neurosci 15: 156-169, 2000.
108
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
2956 448. Zundel W, and Giaccia A. Inhibition of the anti-apoptotic PI(3)K/Akt/Bad pathway by stress.
2957 Genes Dev 12: 1941-1946, 1998.
2958 449. Kanety H, Hemi R, Papa MZ, and Karasik A. Sphingomyelinase and ceramide suppress insulin-
2959 induced tyrosine phosphorylation of the insulin receptor substrate-1. Journal of Biological Chemistry
2960 271: 9895-9897, 1996.
2961 450. Peraldi P, Hotamisligil GS, Buurman WA, White MF, and Spiegelman BM. Tumor necrosis factor
2962 (TNF)-alpha inhibits insulin signaling through stimulation of the p55 TNF receptor and activation of
2963 sphingomyelinase. J Biol Chem 271: 13018-13022, 1996.
2964 451. Bharath LP, Ruan T, Li Y, Ravindran A, Wan X, Nhan JK, Walker ML, Deeter L, Goodrich R,
2965 Johnson E, Munday D, Mueller R, Kunz D, Jones D, Reese V, Summers SA, Babu PV, Holland WL, Zhang
2966 QJ, Abel ED, and Symons JD. Ceramide-Initiated Protein Phosphatase 2A Activation Contributes to
2967 Arterial Dysfunction In Vivo. Diabetes 64: 3914-3926, 2015.
2968 452. Hoehn KL, Hohnen-Behrens C, Cederberg A, Wu LE, Turner N, Yuasa T, Ebina Y, and James DE.
2969 IRS1-independent defects define major nodes of insulin resistance. Cell Metab 7: 421-433, 2008.
2970 453. Hoehn KL, Salmon AB, Hohnen-Behrens C, Turner N, Hoy AJ, Maghzal GJ, Stocker R, Van
2971 Remmen H, Kraegen EW, Cooney GJ, Richardson AR, and James DE. Insulin resistance is a cellular
2972 antioxidant defense mechanism. Proc Natl Acad Sci U S A 106: 17787-17792, 2009.
2973 454. Houstis N, Rosen ED, and Lander ES. Reactive oxygen species have a causal role in multiple
2974 forms of insulin resistance. Nature 440: 944-948, 2006.
2975 455. Taddeo EP, Laker RC, Breen DS, Akhtar YN, Kenwood BM, Liao JA, Zhang M, Fazakerley DJ,
2976 Tomsig JL, Harris TE, Keller SR, Chow JD, Lynch KR, Chokki M, Molkentin JD, Turner N, James DE, Yan Z,
2977 and Hoehn KL. Opening of the mitochondrial permeability transition pore links mitochondrial
2978 dysfunction to insulin resistance in skeletal muscle. Mol Metab 3: 124-134, 2014.
2979 456. JeBailey L, Wanono O, Niu W, Roessler J, Rudich A, and Klip A. Ceramide- and oxidant-induced
2980 insulin resistance involve loss of insulin-dependent Rac-activation and actin remodeling in muscle cells.
2981 Diabetes 56: 394-403, 2007.
2982 457. Foley KP, and Klip A. Dynamic GLUT4 sorting through a syntaxin-6 compartment in muscle cells
2983 is derailed by insulin resistance-causing ceramide. Biol Open 3: 314-325, 2014.
2984 458. Kim YR, Lee EJ, Shin KO, Kim MH, Pewzner-Jung Y, Lee YM, Park JW, Futerman AH, and Park
2985 WJ. Hepatic triglyceride accumulation via endoplasmic reticulum stress-induced SREBP-1 activation is
2986 regulated by ceramide synthases. Exp Mol Med 51: 1-16, 2019.
2987 459. Chaurasia B, and Summers SA. Ceramides in Metabolism: Key Lipotoxic Players. Annu Rev
2988 Physiol 83: 303-330, 2021.
2989 460. Unger RH. Lipotoxicity in the pathogenesis of obesity-dependent NIDDM. Genetic and clinical
2990 implications. Diabetes 44: 863-870, 1995.
2991 461. Pearson GL, Mellett N, Chu KY, Boslem E, Meikle PJ, and Biden TJ. A comprehensive lipidomic
2992 screen of pancreatic beta-cells using mass spectroscopy defines novel features of glucose-stimulated
2993 turnover of neutral lipids, sphingolipids and plasmalogens. Mol Metab 5: 404-414, 2016.
2994 462. Boslem E, Meikle PJ, and Biden TJ. Roles of ceramide and sphingolipids in pancreatic beta-cell
2995 function and dysfunction. Islets 4: 177-187, 2012.
2996 463. Listenberger LL, Han X, Lewis SE, Cases S, Farese RV, Jr., Ory DS, and Schaffer JE. Triglyceride
2997 accumulation protects against fatty acid-induced lipotoxicity. Proc Natl Acad Sci U S A 100: 3077-3082,
2998 2003.
2999 464. Lang F, Ullrich S, and Gulbins E. Ceramide formation as a target in beta-cell survival and
3000 function. Expert Opin Ther Targets 15: 1061-1071, 2011.
3001 465. Veret J, Coant N, Berdyshev EV, Skobeleva A, Therville N, Bailbe D, Gorshkova I, Natarajan V,
3002 Portha B, and Le Stunff H. Ceramide synthase 4 and de novo production of ceramides with specific N-
109
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3003 acyl chain lengths are involved in glucolipotoxicity-induced apoptosis of INS-1 beta-cells. Biochem J 438:
3004 177-189, 2011.
3005 466. Manukyan L, Ubhayasekera SJ, Bergquist J, Sargsyan E, and Bergsten P. Palmitate-induced
3006 impairments of beta-cell function are linked with generation of specific ceramide species via acylation of
3007 sphingosine. Endocrinology 156: 802-812, 2015.
3008 467. Sjoholm A. Ceramide inhibits pancreatic beta-cell insulin production and mitogenesis and
3009 mimics the actions of interleukin-1 beta. FEBS Lett 367: 283-286, 1995.
3010 468. Boslem E, MacIntosh G, Preston AM, Bartley C, Busch AK, Fuller M, Laybutt DR, Meikle PJ, and
3011 Biden TJ. A lipidomic screen of palmitate-treated MIN6 beta-cells links sphingolipid metabolites with
3012 endoplasmic reticulum (ER) stress and impaired protein trafficking. Biochem J 435: 267-276, 2011.
3013 469. Wehinger S, Ortiz R, Diaz MI, Aguirre A, Valenzuela M, Llanos P, Mc Master C, Leyton L, and
3014 Quest AF. Phosphorylation of caveolin-1 on tyrosine-14 induced by ROS enhances palmitate-induced
3015 death of beta-pancreatic cells. Biochim Biophys Acta 1852: 693-708, 2015.
3016 470. Syed I, Szulc ZM, Ogretmen B, and Kowluru A. L- threo -C6-pyridinium-ceramide bromide, a
3017 novel cationic ceramide, induces NADPH oxidase activation, mitochondrial dysfunction and loss in cell
3018 viability in INS 832/13 beta-cells. Cell Physiol Biochem 30: 1051-1058, 2012.
3019 471. Syed I, Jayaram B, Subasinghe W, and Kowluru A. Tiam1/Rac1 signaling pathway mediates
3020 palmitate-induced, ceramide-sensitive generation of superoxides and lipid peroxides and the loss of
3021 mitochondrial membrane potential in pancreatic beta-cells. Biochem Pharmacol 80: 874-883, 2010.
3022 472. Kowluru A, and Kowluru RA. RACking up ceramide-induced islet beta-cell dysfunction. Biochem
3023 Pharmacol 154: 161-169, 2018.
3024 473. Hali M, Wadzinski BE, and Kowluru A. Alpha4 contributes to the dysfunction of the pancreatic
3025 beta cell under metabolic stress. Mol Cell Endocrinol 557: 111754, 2022.
3026 474. Castell AL, Vivoli A, Tippetts TS, Frayne IR, Angeles ZE, Moulle VS, Campbell SA, Ruiz M,
3027 Ghislain J, Des Rosiers C, Holland WL, Summers SA, and Poitout V. Very-Long-Chain Unsaturated
3028 Sphingolipids Mediate Oleate-Induced Rat beta-Cell Proliferation. Diabetes 71: 1218-1232, 2022.
3029 475. Griess K, Rieck M, Muller N, Karsai G, Hartwig S, Pelligra A, Hardt R, Schlegel C, Kuboth J,
3030 Uhlemeyer C, Trenkamp S, Jeruschke K, Weiss J, Peifer-Weiss L, Xu W, Cames S, Yi X, Cnop M, Beller M,
3031 Stark H, Kondadi AK, Reichert AS, Markgraf D, Wammers M, Haussinger D, Kuss O, Lehr S, Eizirik D,
3032 Lickert H, Lammert E, Roden M, Winter D, Al-Hasani H, Hoglinger D, Hornemann T, Bruning JC, and
3033 Belgardt BF. Sphingolipid subtypes differentially control proinsulin processing and systemic glucose
3034 homeostasis. Nat Cell Biol 25: 20-29, 2023.
3035 476. Raichur S, Wang ST, Chan PW, Li Y, Ching J, Chaurasia B, Dogra S, Öhman MK, Takeda K, Sugii
3036 S, Pewzner-Jung Y, Futerman AH, and Summers SA. CerS2 haploinsufficiency inhibits β-oxidation and
3037 confers susceptibility to diet-induced steatohepatitis and insulin resistance. Cell Metab 20: 687-695,
3038 2014.
3039 477. Pewzner-Jung Y, Park H, Laviad EL, Silva LC, Lahiri S, Stiban J, Erez-Roman R, Brugger B,
3040 Sachsenheimer T, Wieland F, Prieto M, Merrill AH, Jr., and Futerman AH. A critical role for ceramide
3041 synthase 2 in liver homeostasis: I. alterations in lipid metabolic pathways. J Biol Chem 285: 10902-10910,
3042 2010.
3043 478. Fothergill E, Guo J, Howard L, Kerns JC, Knuth ND, Brychta R, Chen KY, Skarulis MC, Walter M,
3044 Walter PJ, and Hall KD. Persistent metabolic adaptation 6 years after "The Biggest Loser" competition.
3045 Obesity (Silver Spring) 24: 1612-1619, 2016.
3046 479. Hammerschmidt P, and Bruning JC. Contribution of specific ceramides to obesity-associated
3047 metabolic diseases. Cell Mol Life Sci 79: 395, 2022.
3048 480. Gao S, Zhu G, Gao X, Wu D, Carrasco P, Casals N, Hegardt FG, Moran TH, and Lopaschuk GD.
3049 Important roles of brain-specific carnitine palmitoyltransferase and ceramide metabolism in leptin
3050 hypothalamic control of feeding. Proc Natl Acad Sci U S A 108: 9691-9696, 2011.
110
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3051 481. Moulle VS, Picard A, Le Foll C, Levin BE, and Magnan C. Lipid sensing in the brain and regulation
3052 of energy balance. Diabetes Metab 40: 29-33, 2014.
3053 482. Borg ML, Omran SF, Weir J, Meikle PJ, and Watt MJ. Consumption of a high-fat diet, but not
3054 regular endurance exercise training, regulates hypothalamic lipid accumulation in mice. J Physiol 590:
3055 4377-4389, 2012.
3056 483. Car H, Zendzian-Piotrowska M, Prokopiuk S, Fiedorowicz A, Sadowska A, Kurek K, and Sawicka
3057 D. Ceramide profiles in the brain of rats with diabetes induced by streptozotocin. FEBS J 279: 1943-1952,
3058 2012.
3059 484. Contreras C, Gonzalez-Garcia I, Martinez-Sanchez N, Seoane-Collazo P, Jacas J, Morgan DA,
3060 Serra D, Gallego R, Gonzalez F, Casals N, Nogueiras R, Rahmouni K, Dieguez C, and Lopez M. Central
3061 ceramide-induced hypothalamic lipotoxicity and ER stress regulate energy balance. Cell Rep 9: 366-377,
3062 2014.
3063 485. Martinez-Sanchez N, Seoane-Collazo P, Contreras C, Varela L, Villarroya J, Rial-Pensado E,
3064 Buque X, Aurrekoetxea I, Delgado TC, Vazquez-Martinez R, Gonzalez-Garcia I, Roa J, Whittle AJ,
3065 Gomez-Santos B, Velagapudi V, Tung YCL, Morgan DA, Voshol PJ, Martinez de Morentin PB, Lopez-
3066 Gonzalez T, Linares-Pose L, Gonzalez F, Chatterjee K, Sobrino T, Medina-Gomez G, Davis RJ, Casals N,
3067 Oresic M, Coll AP, Vidal-Puig A, Mittag J, Tena-Sempere M, Malagon MM, Dieguez C, Martinez-Chantar
3068 ML, Aspichueta P, Rahmouni K, Nogueiras R, Sabio G, Villarroya F, and Lopez M. Hypothalamic AMPK-
3069 ER Stress-JNK1 Axis Mediates the Central Actions of Thyroid Hormones on Energy Balance. Cell Metab
3070 26: 212-229 e212, 2017.
3071 486. Gonzalez-Garcia I, Contreras C, Estevez-Salguero A, Ruiz-Pino F, Colsh B, Pensado I, Linares-
3072 Pose L, Rial-Pensado E, Martinez de Morentin PB, Ferno J, Dieguez C, Nogueiras R, Le Stunff H,
3073 Magnan C, Tena-Sempere M, and Lopez M. Estradiol Regulates Energy Balance by Ameliorating
3074 Hypothalamic Ceramide-Induced ER Stress. Cell Rep 25: 413-423 e415, 2018.
3075 487. Ramirez S, Martins L, Jacas J, Carrasco P, Pozo M, Clotet J, Serra D, Hegardt FG, Dieguez C,
3076 Lopez M, and Casals N. Hypothalamic ceramide levels regulated by CPT1C mediate the orexigenic effect
3077 of ghrelin. Diabetes 62: 2329-2337, 2013.
3078 488. Nordstrom V, Willershauser M, Herzer S, Rozman J, von Bohlen Und Halbach O, Meldner S,
3079 Rothermel U, Kaden S, Roth FC, Waldeck C, Gretz N, de Angelis MH, Draguhn A, Klingenspor M, Grone
3080 HJ, and Jennemann R. Neuronal expression of glucosylceramide synthase in central nervous system
3081 regulates body weight and energy homeostasis. PLoS Biol 11: e1001506, 2013.
3082 489. Herzer S, Meldner S, Grone HJ, and Nordstrom V. Fasting-Induced Lipolysis and Hypothalamic
3083 Insulin Signaling Are Regulated by Neuronal Glucosylceramide Synthase. Diabetes 64: 3363-3376, 2015.
3084 490. Gao Y, Layritz C, Legutko B, Eichmann TO, Laperrousaz E, Moulle VS, Cruciani-Guglielmacci C,
3085 Magnan C, Luquet S, Woods SC, Eckel RH, Yi CX, Garcia-Caceres C, and Tschop MH. Disruption of Lipid
3086 Uptake in Astroglia Exacerbates Diet-Induced Obesity. Diabetes 66: 2555-2563, 2017.
3087 491. Dale N. Purinergic signaling in hypothalamic tanycytes: potential roles in chemosensing. Semin
3088 Cell Dev Biol 22: 237-244, 2011.
3089 492. Alford S, Patel D, Perakakis N, and Mantzoros CS. Obesity as a risk factor for Alzheimer's
3090 disease: weighing the evidence. Obes Rev 19: 269-280, 2018.
3091 493. Ramirez A, Wolfsgruber S, Lange C, Kaduszkiewicz H, Weyerer S, Werle J, Pentzek M, Fuchs A,
3092 Riedel-Heller SG, Luck T, Mosch E, Bickel H, Wiese B, Prokein J, Konig HH, Brettschneider C, Breteler
3093 MM, Maier W, Jessen F, Scherer M, and AgeCoDe Study G. Elevated HbA1c is associated with increased
3094 risk of incident dementia in primary care patients. J Alzheimers Dis 44: 1203-1212, 2015.
3095 494. Kaya I, Zetterberg H, Blennow K, and Hanrieder J. Shedding Light on the Molecular Pathology of
3096 Amyloid Plaques in Transgenic Alzheimer's Disease Mice Using Multimodal MALDI Imaging Mass
3097 Spectrometry. ACS Chem Neurosci 9: 1802-1817, 2018.
111
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3098 495. Strnad S, PraZienkova V, Holubova M, Sykora D, Cvacka J, Maletinska L, Zelezna B, Kunes J,
3099 and Vrkoslav V. Mass spectrometry imaging of free-floating brain sections detects pathological lipid
3100 distribution in a mouse model of Alzheimer's-like pathology. Analyst 145: 4595-4605, 2020.
3101 496. Filippov V, Song MA, Zhang K, Vinters HV, Tung S, Kirsch WM, Yang J, and Duerksen-Hughes PJ.
3102 Increased ceramide in brains with Alzheimer's and other neurodegenerative diseases. J Alzheimers Dis
3103 29: 537-547, 2012.
3104 497. Mielke MM, Bandaru VV, Haughey NJ, Xia J, Fried LP, Yasar S, Albert M, Varma V, Harris G,
3105 Schneider EB, Rabins PV, Bandeen-Roche K, Lyketsos CG, and Carlson MC. Serum ceramides increase
3106 the risk of Alzheimer disease: the Women's Health and Aging Study II. Neurology 79: 633-641, 2012.
3107 498. Dinkins MB, Enasko J, Hernandez C, Wang G, Kong J, Helwa I, Liu Y, Terry AV, Jr., and Bieberich
3108 E. Neutral Sphingomyelinase-2 Deficiency Ameliorates Alzheimer's Disease Pathology and Improves
3109 Cognition in the 5XFAD Mouse. J Neurosci 36: 8653-8667, 2016.
3110 499. Rosen ED, and Spiegelman BM. Adipocytes as regulators of energy balance and glucose
3111 homeostasis. Nature 444: 847-853, 2006.
3112 500. Rosen ED, and Spiegelman BM. What we talk about when we talk about fat. Cell 156: 20-44,
3113 2014.
3114 501. Virtue S, and Vidal-Puig A. Adipose tissue expandability, lipotoxicity and the Metabolic
3115 Syndrome--an allostatic perspective. Biochim Biophys Acta 1801: 338-349, 2010.
3116 502. Blachnio-Zabielska AU, Baranowski M, Hirnle T, Zabielski P, Lewczuk A, Dmitruk I, and Gorski J.
3117 Increased bioactive lipids content in human subcutaneous and epicardial fat tissue correlates with
3118 insulin resistance. Lipids 47: 1131-1141, 2012.
3119 503. Blachnio-Zabielska AU, Koutsari C, Tchkonia T, and Jensen MD. Sphingolipid content of human
3120 adipose tissue: relationship to adiponectin and insulin resistance. Obesity (Silver Spring) 20: 2341-2347,
3121 2012.
3122 504. Choromanska B, Mysliwiec P, Razak Hady H, Dadan J, Mysliwiec H, Chabowski A, and Miklosz
3123 A. Metabolic Syndrome is Associated with Ceramide Accumulation in Visceral Adipose Tissue of Women
3124 with Morbid Obesity. Obesity (Silver Spring) 27: 444-453, 2019.
3125 505. Song Y, Zan W, Qin L, Han S, Ye L, Wang M, Jiang B, Fang P, Liu Q, Shao C, Gong Y, and Li P.
3126 Ablation of ORMDL3 impairs adipose tissue thermogenesis and insulin sensitivity by increasing ceramide
3127 generation. Mol Metab 56: 101423, 2022.
3128 506. Pan DZ, Garske KM, Alvarez M, Bhagat YV, Boocock J, Nikkola E, Miao Z, Raulerson CK, Cantor
3129 RM, Civelek M, Glastonbury CA, Small KS, Boehnke M, Lusis AJ, Sinsheimer JS, Mohlke KL, Laakso M,
3130 Pajukanta P, and Ko A. Integration of human adipocyte chromosomal interactions with adipose gene
3131 expression prioritizes obesity-related genes from GWAS. Nat Commun 9: 1512, 2018.
3132 507. Chaurasia B, Holland WL, and Summers SA. Does This Schlank Make Me Look Fat? Trends
3133 Endocrinol Metab 29: 597-599, 2018.
3134 508. Lee SY, Lee HY, Song JH, Kim GT, Jeon S, Song YJ, Lee JS, Hur JH, Oh HH, Park SY, Shim SM, Yoo
3135 HJ, Lee BC, Jiang XC, Choi CS, and Park TS. Adipocyte-Specific Deficiency of De Novo Sphingolipid
3136 Biosynthesis Leads to Lipodystrophy and Insulin Resistance. Diabetes 66: 2596-2609, 2017.
3137 509. Alexaki A, Clarke BA, Gavrilova O, Ma Y, Zhu H, Ma X, Xu L, Tuymetova G, Larman BC, Allende
3138 ML, Dunn TM, and Proia RL. De Novo Sphingolipid Biosynthesis Is Required for Adipocyte Survival and
3139 Metabolic Homeostasis. J Biol Chem 292: 3929-3939, 2017.
3140 510. Lange M, Angelidou G, Ni Z, Criscuolo A, Schiller J, Bluher M, and Fedorova M. AdipoAtlas: A
3141 reference lipidome for human white adipose tissue. Cell Rep Med 2: 100407, 2021.
3142 511. Raichur S, Brunner B, Bielohuby M, Hansen G, Pfenninger A, Wang B, Bruning JC, Larsen PJ,
3143 and Tennagels N. The role of C16:0 ceramide in the development of obesity and type 2 diabetes: CerS6
3144 inhibition as a novel therapeutic approach. Mol Metab 21: 36-50, 2019.
112
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3145 512. Rawshani A, Rawshani A, Franzen S, Eliasson B, Svensson AM, Miftaraj M, McGuire DK, Sattar
3146 N, Rosengren A, and Gudbjornsdottir S. Mortality and Cardiovascular Disease in Type 1 and Type 2
3147 Diabetes. N Engl J Med 376: 1407-1418, 2017.
3148 513. Strain WD, and Paldanius PM. Diabetes, cardiovascular disease and the microcirculation.
3149 Cardiovasc Diabetol 17: 57, 2018.
3150 514. Glovaci D, Fan W, and Wong ND. Epidemiology of Diabetes Mellitus and Cardiovascular Disease.
3151 Curr Cardiol Rep 21: 21, 2019.
3152 515. Vetrone LM, Zaccardi F, Webb DR, Seidu S, Gholap NN, Pitocco D, Davies MJ, and Khunti K.
3153 Cardiovascular and mortality events in type 2 diabetes cardiovascular outcomes trials: a systematic
3154 review with trend analysis. Acta Diabetol 56: 331-339, 2019.
3155 516. Simmonds SJ, Cuijpers I, Heymans S, and Jones EAV. Cellular and Molecular Differences
3156 between HFpEF and HFrEF: A Step Ahead in an Improved Pathological Understanding. Cells 9: 2020.
3157 517. Dunlay SM, Givertz MM, Aguilar D, Allen LA, Chan M, Desai AS, Deswal A, Dickson VV,
3158 Kosiborod MN, Lekavich CL, McCoy RG, Mentz RJ, and Piña IL. Type 2 Diabetes Mellitus and Heart
3159 Failure: A Scientific Statement From the American Heart Association and the Heart Failure Society of
3160 America: This statement does not represent an update of the 2017 ACC/AHA/HFSA heart failure
3161 guideline update. Circulation 140: e294-e324, 2019.
3162 518. Pogatsa G. Metabolic energy metabolism in diabetes: therapeutic implications. Coron Artery Dis
3163 12 Suppl 1: S29-33, 2001.
3164 519. Mondal LK, Baidya KP, Bhattacharya B, Giri A, and Bhaduri G. Relation between increased
3165 anaerobic glycolysis and visual acuity in long-standing type 2 diabetes mellitus without retinopathy.
3166 Indian J Ophthalmol 54: 43-44, 2006.
3167 520. B W. Ceramides, Plasma [A Test in Focus].
3168 https://news.mayomedicallaboratories.com/2016/07/28/ceramides-plasma-a-test-in-focus/.
3169 521. Laaksonen R, Ekroos K, Sysi-Aho M, Hilvo M, Vihervaara T, Kauhanen D, Suoniemi M, Hurme R,
3170 März W, Scharnagl H, Stojakovic T, Vlachopoulou E, Lokki ML, Nieminen MS, Klingenberg R, Matter
3171 CM, Hornemann T, Jüni P, Rodondi N, Räber L, Windecker S, Gencer B, Pedersen ER, Tell GS, Nygård O,
3172 Mach F, Sinisalo J, and Lüscher TF. Plasma ceramides predict cardiovascular death in patients with
3173 stable coronary artery disease and acute coronary syndromes beyond LDL-cholesterol. Eur Heart J 37:
3174 1967-1976, 2016.
3175 522. Hilvo M, Meikle PJ, Pedersen ER, Tell GS, Dhar I, Brenner H, Schottker B, Laaperi M, Kauhanen
3176 D, Koistinen KM, Jylha A, Huynh K, Mellett NA, Tonkin AM, Sullivan DR, Simes J, Nestel P, Koenig W,
3177 Rothenbacher D, Nygard O, and Laaksonen R. Development and validation of a ceramide- and
3178 phospholipid-based cardiovascular risk estimation score for coronary artery disease patients. Eur Heart J
3179 2019.
3180 523. Poss AM, Maschek JA, Cox JE, Hauner BJ, Hopkins PN, Hunt SC, Holland WL, Summers SA, and
3181 Playdon MC. Machine learning reveals serum sphingolipids as cholesterol-independent biomarkers of
3182 coronary artery disease. J Clin Invest 130: 1363-1376, 2020.
3183 524. Havulinna AS, Sysi-Aho M, Hilvo M, Kauhanen D, Hurme R, Ekroos K, Salomaa V, and
3184 Laaksonen R. Circulating Ceramides Predict Cardiovascular Outcomes in the Population-Based FINRISK
3185 2002 Cohort. Arterioscler Thromb Vasc Biol 36: 2424-2430, 2016.
3186 525. Mantovani A, Bonapace S, Lunardi G, Salgarello M, Dugo C, Canali G, Byrne CD, Gori S, Barbieri
3187 E, and Targher G. Association between plasma ceramides and inducible myocardial ischemia in patients
3188 with established or suspected coronary artery disease undergoing myocardial perfusion scintigraphy.
3189 Metabolism 85: 305-312, 2018.
3190 526. Mantovani A, Bonapace S, Lunardi G, Salgarello M, Dugo C, Gori S, Barbieri E, Verlato G,
3191 Laaksonen R, Byrne CD, and Targher G. Association of Plasma Ceramides With Myocardial Perfusion in
113
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3192 Patients With Coronary Artery Disease Undergoing Stress Myocardial Perfusion Scintigraphy. Arterioscler
3193 Thromb Vasc Biol 38: 2854-2861, 2018.
3194 527. Leiherer A, Muendlein A, Saely CH, Geiger K, Brandtner EM, Heinzle C, Gaenger S, Mink S,
3195 Laaksonen R, Fraunberger P, and Drexel H. Coronary Event Risk Test (CERT) as a Risk Predictor for the
3196 10-Year Clinical Outcome of Patients with Peripheral Artery Disease. J Clin Med 12: 2023.
3197 528. Peterson LR, Xanthakis V, Duncan MS, Gross S, Friedrich N, Volzke H, Felix SB, Jiang H, Sidhu R,
3198 Nauck M, Jiang X, Ory DS, Dorr M, Vasan RS, and Schaffer JE. Ceramide Remodeling and Risk of
3199 Cardiovascular Events and Mortality. J Am Heart Assoc 7: 2018.
3200 529. Cresci S, Zhang R, Yang Q, Duncan MS, Xanthakis V, Jiang X, Vasan RS, Schaffer JE, and
3201 Peterson LR. Genetic Architecture of Circulating Very-Long-Chain (C24:0 and C22:0) Ceramide
3202 Concentrations. J Lipid Atheroscler 9: 172-183, 2020.
3203 530. Lemaitre RN, Jensen PN, Hoofnagle A, McKnight B, Fretts AM, King IB, Siscovick DS, Psaty BM,
3204 Heckbert SR, Mozaffarian D, and Sotoodehnia N. Plasma Ceramides and Sphingomyelins in Relation to
3205 Heart Failure Risk. Circ Heart Fail 12: e005708, 2019.
3206 531. Lemaitre RN, Yu C, Hoofnagle A, Hari N, Jensen P, Fretts AM, Umans JG, Howard BV, Sitlani
3207 CM, Siscovick DS, King IB, Sotoodehnia N, and McKnight B. Circulating Sphingolipids, Insulin, HOMA-IR
3208 and HOMA-B: the Strong Heart Family Study. Diabetes 2018.
3209 532. Jensen PN, Fretts AM, Yu C, Hoofnagle AN, Umans JG, Howard BV, Sitlani CM, Siscovick DS,
3210 King IB, Sotoodehnia N, McKnight B, and Lemaitre RN. Circulating sphingolipids, fasting glucose, and
3211 impaired fasting glucose: The Strong Heart Family Study. EBioMedicine 41: 44-49, 2019.
3212 533. Fretts AM, Jensen PN, Hoofnagle A, McKnight B, Howard BV, Umans J, Yu C, Sitlani C, Siscovick
3213 DS, King IB, Sotoodehnia N, and Lemaitre RN. Plasma Ceramide Species Are Associated with Diabetes
3214 Risk in Participants of the Strong Heart Study. J Nutr 150: 1214-1222, 2020.
3215 534. Jensen PN, Fretts AM, Hoofnagle AN, McKnight B, Howard BV, Umans JG, Sitlani CM, Siscovick
3216 DS, King IB, Sotoodehnia N, and Lemaitre RN. Circulating ceramides and sphingomyelins and the risk of
3217 incident cardiovascular disease among people with diabetes: the strong heart study. Cardiovasc Diabetol
3218 21: 167, 2022.
3219 535. Chokshi A, Drosatos K, Cheema FH, Ji R, Khawaja T, Yu S, Kato T, Khan R, Takayama H, Knoll R,
3220 Milting H, Chung CS, Jorde U, Naka Y, Mancini DM, Goldberg IJ, and Schulze PC. Ventricular assist
3221 device implantation corrects myocardial lipotoxicity, reverses insulin resistance, and normalizes cardiac
3222 metabolism in patients with advanced heart failure. Circulation 125: 2844-2853, 2012.
3223 536. Kato TS, Chokshi A, Singh P, Khawaja T, Cheema F, Akashi H, Shahzad K, Iwata S, Homma S,
3224 Takayama H, Naka Y, Jorde U, Farr M, Mancini DM, and Schulze PC. Effects of continuous-flow versus
3225 pulsatile-flow left ventricular assist devices on myocardial unloading and remodeling. Circ Heart Fail 4:
3226 546-553, 2011.
3227 537. Khan RS, Kato TS, Chokshi A, Chew M, Yu S, Wu C, Singh P, Cheema FH, Takayama H, Harris C,
3228 Reyes-Soffer G, Knöll R, Milting H, Naka Y, Mancini D, and Schulze PC. Adipose tissue inflammation and
3229 adiponectin resistance in patients with advanced heart failure: correction after ventricular assist device
3230 implantation. Circ Heart Fail 5: 340-348, 2012.
3231 538. Tsao CW, Lyass A, Enserro D, Larson MG, Ho JE, Kizer JR, Gottdiener JS, Psaty BM, and Vasan
3232 RS. Temporal Trends in the Incidence of and Mortality Associated With Heart Failure With Preserved and
3233 Reduced Ejection Fraction. JACC Heart Fail 6: 678-685, 2018.
3234 539. van Heerebeek L, and Paulus WJ. Understanding heart failure with preserved ejection fraction:
3235 where are we today? Neth Heart J 24: 227-236, 2016.
3236 540. Oktay AA, Rich JD, and Shah SJ. The emerging epidemic of heart failure with preserved ejection
3237 fraction. Curr Heart Fail Rep 10: 401-410, 2013.
3238 541. Unger RH. Lipotoxic Diseases. Annual Review of Medicine 53: 319-336, 2002.
114
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3239 542. Fucho R, Casals N, Serra D, and Herrero L. Ceramides and mitochondrial fatty acid oxidation in
3240 obesity. The FASEB Journal 31: 1263-1272, 2017.
3241 543. Wang J, Zhen L, Klug MG, Wood D, Wu X, and Mizrahi J. Involvement of caspase 3- and 8-like
3242 proteases in ceramide-induced apoptosis of cardiomyocytes. J Card Fail 6: 243-249, 2000.
3243 544. Kruger-Genge A, Blocki A, Franke RP, and Jung F. Vascular Endothelial Cell Biology: An Update.
3244 Int J Mol Sci 20: 2019.
3245 545. McMullan DM, Wilansky S, and Frazier OH. Ventricular fibrillation in a heterotopic heart
3246 transplant recipient. Ann Thorac Surg 75: 598, 2003.
3247 546. Sprague AH, and Khalil RA. Inflammatory cytokines in vascular dysfunction and vascular disease.
3248 Biochem Pharmacol 78: 539-552, 2009.
3249 547. Vallance P, Collier J, and Moncada S. Effects of endothelium-derived nitric oxide on peripheral
3250 arteriolar tone in man. Lancet 2: 997-1000, 1989.
3251 548. Ross R. Atherosclerosis--an inflammatory disease. N Engl J Med 340: 115-126, 1999.
3252 549. Higashi Y, Kihara Y, and Noma K. Endothelial dysfunction and hypertension in aging. Hypertens
3253 Res 35: 1039-1047, 2012.
3254 550. Triggle CR, and Ding H. A review of endothelial dysfunction in diabetes: a focus on the
3255 contribution of a dysfunctional eNOS. J Am Soc Hypertens 4: 102-115, 2010.
3256 551. Wende AR, Symons JD, and Abel ED. Mechanisms of lipotoxicity in the cardiovascular system.
3257 Curr Hypertens Rep 14: 517-531, 2012.
3258 552. Mount PF, Kemp BE, and Power DA. Regulation of endothelial and myocardial NO synthesis by
3259 multi-site eNOS phosphorylation. J Mol Cell Cardiol 42: 271-279, 2007.
3260 553. Smith AR, Visioli F, Frei B, and Hagen TM. Age-related changes in endothelial nitric oxide
3261 synthase phosphorylation and nitric oxide dependent vasodilation: evidence for a novel mechanism
3262 involving sphingomyelinase and ceramide-activated phosphatase 2A. Aging Cell 5: 391-400, 2006.
3263 554. Oaks J, and Ogretmen B. Regulation of PP2A by Sphingolipid Metabolism and Signaling. Front
3264 Oncol 4: 388, 2014.
3265 555. Rajagopalan S, Kurz S, Munzel T, Tarpey M, Freeman BA, Griendling KK, and Harrison DG.
3266 Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane
3267 NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone. The Journal of clinical
3268 investigation 97: 1916-1923, 1996.
3269 556. Hink U, Li H, Mollnau H, Oelze M, Matheis E, Hartmann M, Skatchkov M, Thaiss F, Stahl RA,
3270 Warnholtz A, Meinertz T, Griendling K, Harrison DG, Forstermann U, and Munzel T. Mechanisms
3271 underlying endothelial dysfunction in diabetes mellitus. Circ Res 88: E14-22, 2001.
3272 557. Bryk D, Olejarz W, and Zapolska-Downar D. The role of oxidative stress and NADPH oxidase in
3273 the pathogenesis of atherosclerosis. Postepy Hig Med Dosw (Online) 71: 57-68, 2017.
3274 558. Li H, Junk P, Huwiler A, Burkhardt C, Wallerath T, Pfeilschifter J, and Forstermann U. Dual
3275 effect of ceramide on human endothelial cells: induction of oxidative stress and transcriptional
3276 upregulation of endothelial nitric oxide synthase. Circulation 106: 2250-2256, 2002.
3277 559. Zhang DX, Zou AP, and Li PL. Ceramide-induced activation of NADPH oxidase and endothelial
3278 dysfunction in small coronary arteries. Am J Physiol Heart Circ Physiol 284: H605-612, 2003.
3279 560. Didion SP, and Faraci FM. Ceramide-induced impairment of endothelial function is prevented by
3280 CuZn superoxide dismutase overexpression. Arterioscler Thromb Vasc Biol 25: 90-95, 2005.
3281 561. Funai K, Summers SA, and Rutter J. Reign in the membrane: How common lipids govern
3282 mitochondrial function. Curr Opin Cell Biol 63: 162-173, 2020.
3283 562. Modur V, Zimmerman GA, Prescott SM, and McIntyre TM. Endothelial cell inflammatory
3284 responses to tumor necrosis factor alpha. Ceramide-dependent and -independent mitogen-activated
3285 protein kinase cascades. J Biol Chem 271: 13094-13102, 1996.
115
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3286 563. Xu J, Yeh CH, Chen S, He L, Sensi SL, Canzoniero LM, Choi DW, and Hsu CY. Involvement of de
3287 novo ceramide biosynthesis in tumor necrosis factor-alpha/cycloheximide-induced cerebral endothelial
3288 cell death. J Biol Chem 273: 16521-16526, 1998.
3289 564. Del Gaudio I, Rubinelli L, Sasset L, Wadsack C, Hla T, and Di Lorenzo A. Endothelial Spns2 and
3290 ApoM Regulation of Vascular Tone and Hypertension Via Sphingosine-1-Phosphate. J Am Heart Assoc 10:
3291 e021261, 2021.
3292 565. Jiang XC, Paultre F, Pearson TA, Reed RG, Francis CK, Lin M, Berglund L, and Tall AR. Plasma
3293 sphingomyelin level as a risk factor for coronary artery disease. Arterioscler Thromb Vasc Biol 20: 2614-
3294 2618, 2000.
3295 566. Adachi R, Ogawa K, Matsumoto SI, Satou T, Tanaka Y, Sakamoto J, Nakahata T, Okamoto R,
3296 Kamaura M, and Kawamoto T. Discovery and characterization of selective human sphingomyelin
3297 synthase 2 inhibitors. Eur J Med Chem 136: 283-293, 2017.
3298 567. Wang X, Dong J, Zhao Y, Li Y, and Wu M. Adenovirus-mediated sphingomyelin synthase 2
3299 increases atherosclerotic lesions in ApoE KO mice. Lipids Health Dis 10: 7, 2011.
3300 568. Zhao YR, Dong JB, Li Y, and Wu MP. Sphingomyelin synthase 2 over-expression induces
3301 expression of aortic inflammatory biomarkers and decreases circulating EPCs in ApoE KO mice. Life Sci
3302 90: 867-873, 2012.
3303 569. Liu J, Huan C, Chakraborty M, Zhang H, Lu D, Kuo MS, Cao G, and Jiang XC. Macrophage
3304 sphingomyelin synthase 2 deficiency decreases atherosclerosis in mice. Circ Res 105: 295-303, 2009.
3305 570. Liu J, Zhang H, Li Z, Hailemariam TK, Chakraborty M, Jiang K, Qiu D, Bui HH, Peake DA, Kuo MS,
3306 Wadgaonkar R, Cao G, and Jiang XC. Sphingomyelin synthase 2 is one of the determinants for plasma
3307 and liver sphingomyelin levels in mice. Arterioscler Thromb Vasc Biol 29: 850-856, 2009.
3308 571. Fan Y, Shi F, Liu J, Dong J, Bui HH, Peake DA, Kuo MS, Cao G, and Jiang XC. Selective reduction
3309 in the sphingomyelin content of atherogenic lipoproteins inhibits their retention in murine aortas and
3310 the subsequent development of atherosclerosis. Arterioscler Thromb Vasc Biol 30: 2114-2120, 2010.
3311 572. Li Z, Fan Y, Liu J, Li Y, Huan C, Bui HH, Kuo MS, Park TS, Cao G, and Jiang XC. Impact of
3312 sphingomyelin synthase 1 deficiency on sphingolipid metabolism and atherosclerosis in mice.
3313 Arterioscler Thromb Vasc Biol 32: 1577-1584, 2012.
3314 573. Yano M, Yamamoto T, Nishimura N, Gotoh T, Watanabe K, Ikeda K, Garan Y, Taguchi R, Node
3315 K, Okazaki T, and Oike Y. Increased oxidative stress impairs adipose tissue function in sphingomyelin
3316 synthase 1 null mice. PLoS One 8: e61380, 2013.
3317 574. Schissel SL, Tweedie-Hardman J, Rapp JH, Graham G, Williams KJ, and Tabas I. Rabbit aorta and
3318 human atherosclerotic lesions hydrolyze the sphingomyelin of retained low-density lipoprotein.
3319 Proposed role for arterial-wall sphingomyelinase in subendothelial retention and aggregation of
3320 atherogenic lipoproteins. J Clin Invest 98: 1455-1464, 1996.
3321 575. Ichi I, Nakahara K, Kiso K, and Kojo S. Effect of dietary cholesterol and high fat on ceramide
3322 concentration in rat tissues. Nutrition 23: 570-574, 2007.
3323 576. Cheng JM, Suoniemi M, Kardys I, Vihervaara T, de Boer SP, Akkerhuis KM, Sysi-Aho M, Ekroos
3324 K, Garcia-Garcia HM, Oemrawsingh RM, Regar E, Koenig W, Serruys PW, van Geuns RJ, Boersma E, and
3325 Laaksonen R. Plasma concentrations of molecular lipid species in relation to coronary plaque
3326 characteristics and cardiovascular outcome: Results of the ATHEROREMO-IVUS study. Atherosclerosis
3327 243: 560-566, 2015.
3328 577. Meeusen JW, Donato LJ, Bryant SC, Baudhuin LM, Berger PB, and Jaffe AS. Plasma Ceramides.
3329 Arterioscler Thromb Vasc Biol 38: 1933-1939, 2018.
3330 578. Park TS, Panek RL, Mueller SB, Hanselman JC, Rosebury WS, Robertson AW, Kindt EK, Homan
3331 R, Karathanasis SK, and Rekhter MD. Inhibition of sphingomyelin synthesis reduces atherogenesis in
3332 apolipoprotein E-knockout mice. Circulation 110: 3465-3471, 2004.
116
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3333 579. Li Z, Park TS, Li Y, Pan X, Iqbal J, Lu D, Tang W, Yu L, Goldberg IJ, Hussain MM, and Jiang XC.
3334 Serine palmitoyltransferase (SPT) deficient mice absorb less cholesterol. Biochim Biophys Acta 1791:
3335 297-306, 2009.
3336 580. Mody N, and McIlroy GD. The mechanisms of Fenretinide-mediated anti-cancer activity and
3337 prevention of obesity and type-2 diabetes. Biochem Pharmacol 91: 277-286, 2014.
3338 581. Koh IU, Jun HS, Choi JS, Lim JH, Kim WH, Yoon JB, and Song J. Fenretinide ameliorates insulin
3339 resistance and fatty liver in obese mice. Biol Pharm Bull 35: 369-375, 2012.
3340 582. Busnelli M, Manzini S, Bonacina F, Soldati S, Barbieri SS, Amadio P, Sandrini L, Arnaboldi F,
3341 Donetti E, Laaksonen R, Paltrinieri S, Scanziani E, and Chiesa G. Fenretinide treatment accelerates
3342 atherosclerosis development in apoE-deficient mice in spite of beneficial metabolic effects. Br J
3343 Pharmacol 177: 328-345, 2020.
3344 583. Rinella ME, Lazarus JV, Ratziu V, Francque SM, Sanyal AJ, Kanwal F, Romero D, Abdelmalek
3345 MF, Anstee QM, Arab JP, Arrese M, Bataller R, Beuers U, Boursier J, Bugianesi E, Byrne C, Castro Narro
3346 GE, Chowdhury A, Cortez-Pinto H, Cryer D, Cusi K, El-Kassas M, Klein S, Eskridge W, Fan J, Gawrieh S,
3347 Guy CD, Harrison SA, Kim SU, Koot B, Korenjak M, Kowdley K, Lacaille F, Loomba R, Mitchell-Thain R,
3348 Morgan TR, Powell E, Roden M, Romero-Gomez M, Silva M, Singh SP, Sookoian SC, Spearman CW,
3349 Tiniakos D, Valenti L, Vos MB, Wong VW, Xanthakos S, Yilmaz Y, Younossi Z, Hobbs A, Villota-Rivas M,
3350 Newsome PN, and group NNc. A multi-society Delphi consensus statement on new fatty liver disease
3351 nomenclature. Hepatology 2023.
3352 584. Badmus OO, Hinds TD, Jr., and Stec DE. Mechanisms Linking Metabolic-Associated Fatty Liver
3353 Disease (MAFLD) to Cardiovascular Disease. Curr Hypertens Rep 25: 151-162, 2023.
3354 585. Karim MA, Singal AG, Kum HC, Lee YT, Park S, Rich NE, Noureddin M, and Yang JD. Clinical
3355 Characteristics and Outcomes of Nonalcoholic Fatty Liver Disease-Associated Hepatocellular Carcinoma
3356 in the United States. Clin Gastroenterol Hepatol 2022.
3357 586. Luukkonen PK, Zhou Y, Sadevirta S, Leivonen M, Arola J, Oresic M, Hyotylainen T, and Yki-
3358 Jarvinen H. Ceramides Dissociate Steatosis and Insulin Resistance in the Human Liver in Non-Alcoholic
3359 Fatty Liver Disease. J Hepatol 2016.
3360 587. Luukkonen PK, Zhou Y, Sädevirta S, Leivonen M, Arola J, Orešič M, Hyötyläinen T, and Yki-
3361 Järvinen H. Hepatic ceramides dissociate steatosis and insulin resistance in patients with non-alcoholic
3362 fatty liver disease. J Hepatol 64: 1167-1175, 2016.
3363 588. Simon J, Ouro A, Ala-Ibanibo L, Presa N, Delgado TC, and Martinez-Chantar ML. Sphingolipids
3364 in Non-Alcoholic Fatty Liver Disease and Hepatocellular Carcinoma: Ceramide Turnover. Int J Mol Sci 21:
3365 2019.
3366 589. Masoodi M, Gastaldelli A, Hyotylainen T, Arretxe E, Alonso C, Gaggini M, Brosnan J, Anstee
3367 QM, Millet O, Ortiz P, Mato JM, Dufour JF, and Oresic M. Metabolomics and lipidomics in NAFLD:
3368 biomarkers and non-invasive diagnostic tests. Nat Rev Gastroenterol Hepatol 18: 835-856, 2021.
3369 590. Barr J, Caballeria J, Martinez-Arranz I, Dominguez-Diez A, Alonso C, Muntane J, Perez-
3370 Cormenzana M, Garcia-Monzon C, Mayo R, Martin-Duce A, Romero-Gomez M, Lo Iacono O, Tordjman
3371 J, Andrade RJ, Perez-Carreras M, Le Marchand-Brustel Y, Tran A, Fernandez-Escalante C, Arevalo E,
3372 Garcia-Unzueta M, Clement K, Crespo J, Gual P, Gomez-Fleitas M, Martinez-Chantar ML, Castro A, Lu
3373 SC, Vazquez-Chantada M, and Mato JM. Obesity-dependent metabolic signatures associated with
3374 nonalcoholic fatty liver disease progression. J Proteome Res 11: 2521-2532, 2012.
3375 591. Wasilewska N, Bobrus-Chociej A, Harasim-Symbor E, Tarasow E, Wojtkowska M, Chabowski A,
3376 and Lebensztejn DM. Increased serum concentration of ceramides in obese children with nonalcoholic
3377 fatty liver disease. Lipids Health Dis 17: 216, 2018.
3378 592. Promrat K, Longato L, Wands JR, and de la Monte SM. Weight loss amelioration of non-
3379 alcoholic steatohepatitis linked to shifts in hepatic ceramide expression and serum ceramide levels.
3380 Hepatol Res 41: 754-762, 2011.
117
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3381 593. McGlinchey AJ, Govaere O, Geng D, Ratziu V, Allison M, Bousier J, Petta S, de Oliviera C,
3382 Bugianesi E, Schattenberg JM, Daly AK, Hyotylainen T, Anstee QM, and Oresic M. Metabolic signatures
3383 across the full spectrum of non-alcoholic fatty liver disease. JHEP Rep 4: 100477, 2022.
3384 594. Grammatikos G, Schoell N, Ferreirós N, Bon D, Herrmann E, Farnik H, Köberle V, Piiper A,
3385 Zeuzem S, Kronenberger B, Waidmann O, and Pfeilschifter J. Serum sphingolipidomic analyses reveal
3386 an upregulation of C16- ceramide and sphingosine-1-phosphate in hepatocellular carcinoma. Oncotarget
3387 7: 2016.
3388 595. Wang X, Zhang A, and Sun H. Power of metabolomics in diagnosis and biomarker discovery of
3389 hepatocellular carcinoma. Hepatology 57: 2072-2077, 2013.
3390 596. Alonso C, Fernandez-Ramos D, Varela-Rey M, Martinez-Arranz I, Navasa N, Van Liempd SM,
3391 Lavin Trueba JL, Mayo R, Ilisso CP, de Juan VG, Iruarrizaga-Lejarreta M, delaCruz-Villar L, Minchole I,
3392 Robinson A, Crespo J, Martin-Duce A, Romero-Gomez M, Sann H, Platon J, Van Eyk J, Aspichueta P,
3393 Noureddin M, Falcon-Perez JM, Anguita J, Aransay AM, Martinez-Chantar ML, Lu SC, and Mato JM.
3394 Metabolomic Identification of Subtypes of Nonalcoholic Steatohepatitis. Gastroenterology 152: 1449-
3395 1461 e1447, 2017.
3396 597. Simons N, Isaacs A, Koek GH, Kuc S, Schaper NC, and Brouwers M. PNPLA3, TM6SF2, and
3397 MBOAT7 Genotypes and Coronary Artery Disease. Gastroenterology 152: 912-913, 2017.
3398 598. Chen LZ, Xin YN, Geng N, Jiang M, Zhang DD, and Xuan SY. PNPLA3 I148M variant in
3399 nonalcoholic fatty liver disease: demographic and ethnic characteristics and the role of the variant in
3400 nonalcoholic fatty liver fibrosis. World J Gastroenterol 21: 794-802, 2015.
3401 599. BasuRay S, Wang Y, Smagris E, Cohen JC, and Hobbs HH. Accumulation of PNPLA3 on lipid
3402 droplets is the basis of associated hepatic steatosis. Proc Natl Acad Sci U S A 116: 9521-9526, 2019.
3403 600. Dulai PS, Singh S, Patel J, Soni M, Prokop LJ, Younossi Z, Sebastiani G, Ekstedt M, Hagstrom H,
3404 Nasr P, Stal P, Wong VW, Kechagias S, Hultcrantz R, and Loomba R. Increased risk of mortality by
3405 fibrosis stage in nonalcoholic fatty liver disease: Systematic review and meta-analysis. Hepatology 65:
3406 1557-1565, 2017.
3407 601. Tesfay M, Goldkamp WJ, and Neuschwander-Tetri BA. NASH: The Emerging Most Common
3408 Form of Chronic Liver Disease. Mo Med 115: 225-229, 2018.
3409 602. Li J, Zou B, Yeo YH, Feng Y, Xie X, Lee DH, Fujii H, Wu Y, Kam LY, Ji F, Li X, Chien N, Wei M,
3410 Ogawa E, Zhao C, Wu X, Stave CD, Henry L, Barnett S, Takahashi H, Furusyo N, Eguchi Y, Hsu YC, Lee
3411 TY, Ren W, Qin C, Jun DW, Toyoda H, Wong VW, Cheung R, Zhu Q, and Nguyen MH. Prevalence,
3412 incidence, and outcome of non-alcoholic fatty liver disease in Asia, 1999-2019: a systematic review and
3413 meta-analysis. Lancet Gastroenterol Hepatol 4: 389-398, 2019.
3414 603. Ciardullo S, and Perseghin G. Trends in prevalence of probable fibrotic non-alcoholic
3415 steatohepatitis in the United States, 1999-2016. Liver Int 43: 340-344, 2023.
3416 604. Lee HW, Kim KJ, Jung KS, Chon YE, Huh JH, Park KH, Chung JB, Kim CO, Han KH, and Park JY.
3417 The relationship between visceral obesity and hepatic steatosis measured by controlled attenuation
3418 parameter. PLoS One 12: e0187066, 2017.
3419 605. Manco M. Insulin Resistance and NAFLD: A Dangerous Liaison beyond the Genetics. Children
3420 (Basel) 4: 2017.
3421 606. Kasumov T, Li L, Li M, Gulshan K, Kirwan JP, Liu X, Previs S, Willard B, Smith JD, and
3422 McCullough A. Ceramide as a mediator of non-alcoholic Fatty liver disease and associated
3423 atherosclerosis. PLoS One 10: e0126910, 2015.
3424 607. Kurek K, Piotrowska DM, Wiesiolek-Kurek P, Lukaszuk B, Chabowski A, Gorski J, and Zendzian-
3425 Piotrowska M. Inhibition of ceramide de novo synthesis reduces liver lipid accumulation in rats with
3426 nonalcoholic fatty liver disease. Liver Int 34: 1074-1083, 2014.
3427 608. Holland WL, Xia JY, Johnson JA, Sun K, Pearson MJ, Sharma AX, Quittner-Strom E, Tippetts TS,
3428 Gordillo R, and Scherer PE. Inducible overexpression of adiponectin receptors highlight the roles of
118
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3429 adiponectin-induced ceramidase signaling in lipid and glucose homeostasis. Mol Metab 6: 267-275,
3430 2017.
3431 609. Chaurasia B, Tippetts TS, Monibas RM, Liu J, Li Y, Wang L, Wilkerson JL, Sweeney CR, Pereira
3432 RF, Sumida DH, Maschek JA, Cox JE, Kaddai V, Lancaster GI, Siddique MM, Poss A, Pearson M, Satapati
3433 S, Zhou H, McLaren DG, Previs SF, Chen Y, Qian Y, Petrov A, Wu M, Shen X, Yao J, Nunes CN, Howard
3434 AD, Wang L, Erion MD, Rutter J, Holland WL, Kelley DE, and Summers SA. Targeting a ceramide double
3435 bond improves insulin resistance and hepatic steatosis. Science 2019.
3436 610. Chen B, Sun L, Zeng G, Shen Z, Wang K, Yin L, Xu F, Wang P, Ding Y, Nie Q, Wu Q, Zhang Z, Xia J,
3437 Lin J, Luo Y, Cai J, Krausz KW, Zheng R, Xue Y, Zheng MH, Li Y, Yu C, Gonzalez FJ, and Jiang C. Gut
3438 bacteria alleviate smoking-related NASH by degrading gut nicotine. Nature 610: 562-568, 2022.
3439 611. Li Y, Summers SA, and Holland WL. Gutting out Myc to decrease ceramides. Nat Metab 3: 890-
3440 891, 2021.
3441 612. Li Z, Kabir I, Tietelman G, Huan C, Fan J, Worgall T, and Jiang X-C. Sphingolipid de novo
3442 biosynthesis is essential for intestine cell survival and barrier function. Cell Death & Disease 9: 173, 2018.
3443 613. Ginkel C, Hartmann D, vom Dorp K, Zlomuzica A, Farwanah H, Eckhardt M, Sandhoff R, Degen
3444 J, Rabionet M, Dere E, Dormann P, Sandhoff K, and Willecke K. Ablation of neuronal ceramide synthase
3445 1 in mice decreases ganglioside levels and expression of myelin-associated glycoprotein in
3446 oligodendrocytes. J Biol Chem 287: 41888-41902, 2012.
3447 614. Imgrund S, Hartmann D, Farwanah H, Eckhardt M, Sandhoff R, Degen J, Gieselmann V,
3448 Sandhoff K, and Willecke K. Adult ceramide synthase 2 (CERS2)-deficient mice exhibit myelin sheath
3449 defects, cerebellar degeneration, and hepatocarcinomas. J Biol Chem 284: 33549-33560, 2009.
3450 615. Park JW, Park WJ, Kuperman Y, Boura-Halfon S, Pewzner-Jung Y, and Futerman AH. Ablation of
3451 very long acyl chain sphingolipids causes hepatic insulin resistance in mice due to altered detergent-
3452 resistant membranes. Hepatology 57: 525-532, 2013.
3453 616. Raichur S, Wang ST, Chan PW, Li Y, Ching J, Chaurasia B, Dogra S, Öhman MK, Takeda K, and
3454 Sugii S. CerS2 haploinsufficiency inhibits β-oxidation and confers susceptibility to diet-induced
3455 steatohepatitis and insulin resistance. Cell metabolism 20: 687-695, 2014.
3456 617. Law BA, Liao X, Moore KS, Southard A, Roddy P, Ji R, Szulc Z, Bielawska A, Schulze PC, and
3457 Cowart LA. Lipotoxic very-long-chain ceramides cause mitochondrial dysfunction, oxidative stress, and
3458 cell death in cardiomyocytes. FASEB J 32: 1403-1416, 2018.
3459 618. Eckl K-M, Tidhar R, Thiele H, Oji V, Hausser I, Brodesser S, Preil M-L, Önal-Akan A, Stock F,
3460 Müller D, Becker K, Casper R, Nürnberg G, Altmüller J, Nürnberg P, Traupe H, Futerman AH, and
3461 Hennies HC. Impaired Epidermal Ceramide Synthesis Causes Autosomal Recessive Congenital Ichthyosis
3462 and Reveals the Importance of Ceramide Acyl Chain Length. Journal of Investigative Dermatology 133:
3463 2202-2211, 2013.
3464 619. Jennemann R, Rabionet M, Gorgas K, Epstein S, Dalpke A, Rothermel U, Bayerle A, van der
3465 Hoeven F, Imgrund S, and Kirsch J. Loss of ceramide synthase 3 causes lethal skin barrier disruption.
3466 Human molecular genetics 21: 586-608, 2012.
3467 620. Ebel P, Imgrund S, Vom Dorp K, Hofmann K, Maier H, Drake H, Degen J, Dormann P, Eckhardt
3468 M, Franz T, and Willecke K. Ceramide synthase 4 deficiency in mice causes lipid alterations in sebum
3469 and results in alopecia. Biochem J 461: 147-158, 2014.
3470 621. Peters F, Vorhagen S, Brodesser S, Jakobshagen K, Brüning JC, Niessen CM, and Krönke M.
3471 Ceramide Synthase 4 Regulates Stem Cell Homeostasis and Hair Follicle Cycling. Journal of Investigative
3472 Dermatology 135: 1501-1509, 2015.
3473 622. Roszczyc-Owsiejczuk K, and Zabielski P. Sphingolipids as a Culprit of Mitochondrial Dysfunction
3474 in Insulin Resistance and Type 2 Diabetes. Front Endocrinol (Lausanne) 12: 635175, 2021.
3475 623. Hammerschmidt P, Ostkotte D, Nolte H, Gerl MJ, Jais A, Brunner HL, Sprenger H-G, Awazawa
3476 M, Nicholls HT, Turpin-Nolan SM, Langer T, Krüger M, Brügger B, and Brüning JC. CerS6-Derived
119
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3477 Sphingolipids Interact with Mff and Promote Mitochondrial Fragmentation in Obesity. Cell 177: 1536-
3478 1552.e1523, 2019.
3479 624. Turpin Sarah M, Nicholls Hayley T, Willmes Diana M, Mourier A, Brodesser S, Wunderlich
3480 Claudia M, Mauer J, Xu E, Hammerschmidt P, Brönneke Hella S, Trifunovic A, LoSasso G, Wunderlich
3481 FT, Kornfeld J-W, Blüher M, Krönke M, and Brüning Jens C. Obesity-Induced CerS6-Dependent C16:0
3482 Ceramide Production Promotes Weight Gain and Glucose Intolerance. Cell Metabolism 20: 678-686,
3483 2014.
3484 625. Sikora J, Dworski S, Jones EE, Kamani MA, Micsenyi MC, Sawada T, Le Faouder P, Bertrand-
3485 Michel J, Dupuy A, Dunn CK, Xuan ICY, Casas J, Fabrias G, Hampson DR, Levade T, Drake RR, Medin JA,
3486 and Walkley SU. Acid Ceramidase Deficiency in Mice Results in a Broad Range of Central Nervous
3487 System Abnormalities. The American Journal of Pathology 187: 864-883, 2017.
3488 626. Bhat OM, Li G, Yuan X, Huang D, Gulbins E, Kukreja RC, and Li P-L. Arterial Medial Calcification
3489 through Enhanced small Extracellular Vesicle Release in Smooth Muscle-Specific Asah1 Gene Knockout
3490 Mice. Scientific Reports 10: 1645, 2020.
3491 627. Correnti J, Lin C, Brettschneider J, Kuriakose A, Jeon S, Scorletti E, Oranu A, McIver-Jenkins D,
3492 Kaneza I, Buyco D, Saiman Y, Furth EE, Argemi J, Bataller R, Holland WL, and Carr RM. Liver-specific
3493 ceramide reduction alleviates steatosis and insulin resistance in alcohol-fed mice. J Lipid Res 61: 983-
3494 994, 2020.
3495 628. Xia JY, Holland WL, Kusminski CM, Sun K, Sharma AX, Pearson MJ, Sifuentes AJ, McDonald JG,
3496 Gordillo R, and Scherer PE. Targeted induction of ceramide degradation leads to improved systemic
3497 metabolism and reduced hepatic steatosis. Cell metabolism 22: 266-278, 2015.
3498 629. Kono M, Dreier JL, Ellis JM, Allende ML, Kalkofen DN, Sanders KM, Bielawski J, Bielawska A,
3499 Hannun YA, and Proia RL. Neutral Ceramidase Encoded by the <em>Asah2</em> Gene Is Essential for
3500 the Intestinal Degradation of Sphingolipids *. Journal of Biological Chemistry 281: 7324-7331, 2006.
3501 630. Seito N, Yamashita T, Tsukuda Y, Matsui Y, Urita A, Onodera T, Mizutani T, Haga H, Fujitani N,
3502 Shinohara Y, Minami A, and Iwasaki N. Interruption of glycosphingolipid synthesis enhances
3503 osteoarthritis development in mice. Arthritis & Rheumatism 64: 2579-2588, 2012.
3504 631. Nordström V, Willershäuser M, Herzer S, Rozman J, von Bohlen und Halbach O, Meldner S,
3505 Rothermel U, Kaden S, Roth FC, and Waldeck C. Neuronal expression of glucosylceramide synthase in
3506 central nervous system regulates body weight and energy homeostasis. PLoS biology 11: e1001506,
3507 2013.
3508 632. Herzer S, Meldner S, Gröne H-J, and Nordström V. Fasting-induced lipolysis and hypothalamic
3509 insulin signaling are regulated by neuronal glucosylceramide synthase. Diabetes 64: 3363-3376, 2015.
3510 633. Jennemann R, Kaden S, Sandhoff R, Nordström V, Wang S, Volz M, Robine S, Amen N,
3511 Rothermel U, and Wiegandt H. Glycosphingolipids are essential for intestinal endocytic function. Journal
3512 of Biological Chemistry 287: 32598-32616, 2012.
3513 634. Gabandé-Rodríguez E, Boya P, Labrador V, Dotti CG, and Ledesma MD. High sphingomyelin
3514 levels induce lysosomal damage and autophagy dysfunction in Niemann Pick disease type A. Cell Death
3515 & Differentiation 21: 864-875, 2014.
3516 635. Macauley SL, Sidman RL, Schuchman EH, Taksir T, and Stewart GR. Neuropathology of the acid
3517 sphingomyelinase knockout mouse model of Niemann-Pick A disease including structure-function
3518 studies associated with cerebellar Purkinje cell degeneration. Exp Neurol 214: 181-192, 2008.
3519 636. Hua G, and Kolesnick R. Using ASMase knockout mice to model human diseases. Handb Exp
3520 Pharmacol 216: 29-54, 2013.
3521 637. Hose M, Günther A, Abberger H, Begum S, Korencak M, Becker KA, Buer J, Westendorf AM,
3522 and Hansen W. T Cell-Specific Overexpression of Acid Sphingomyelinase Results in Elevated T Cell
3523 Activation and Reduced Parasitemia During Plasmodium yoelii Infection. Frontiers in Immunology 10:
3524 2019.
120
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3525 638. Smith EL, and Schuchman EH. Acid Sphingomyelinase Overexpression Enhances the
3526 Antineoplastic Effects of Irradiation <em>In Vitro</em> and <em>In Vivo</em>. Molecular Therapy 16:
3527 1565-1571, 2008.
3528 639. Stoffel W, Jenke B, Block B, Zumbansen M, and Koebke J. Neutral sphingomyelinase 2 (smpd3)
3529 in the control of postnatal growth and development. Proc Natl Acad Sci U S A 102: 4554-4559, 2005.
3530
3531
121
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.
3532
3533
3534
122
Downloaded from journals.physiology.org/journal/physrev at Univ El Bosque Biblio (144.202.001.126) on March 11, 2024.