Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Cellular Signalling 79 (2021) 109890

Contents lists available at ScienceDirect

Cellular Signalling
journal homepage: www.elsevier.com/locate/cellsig

The noncanonical chronicles: Emerging roles of sphingolipid


structural variants
Brenda Wan Shing Lam a, Ting Yu Amelia Yam a, Christopher P. Chen a, b, Mitchell K.P. Lai a, b,
Wei-Yi Ong c, Deron R. Herr a, d, e, *
a
Department of Pharmacology, Yong Loo Lin School of Medicine, National University of Singapore, Singapore
b
Memory Aging and Cognition Centre, Department of Psychological Medicine, Yong Loo Lin School of Medicine, National University of Singapore, Singapore
c
Department of Anatomy, Yong Loo Lin School of Medicine, National University of Singapore, Singapore
d
Department of Biology, San Diego State University, San Diego, CA, USA
e
American University of Health Sciences, Long Beach, CA, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Sphingolipids (SPs) are structurally diverse and represent one of the most quantitatively abundant classes of
Chain-length variation lipids in mammalian cells. In addition to their structural roles, many SP species are known to be bioactive
Lipidomics mediators of essential cellular processes. Historically, studies have focused on SP species that contain the ca­
Neuroinflammation
nonical 18‑carbon, mono-unsaturated sphingoid backbone. However, increasingly sensitive analytical technol­
Phytosphingosine
Sphingadiene
ogies, driven by advances in mass spectrometry, have facilitated the identification of previously under-
Sphingoid backbone appreciated, molecularly distinct SP species. Many of these less abundant species contain noncanonical back­
bones. Interestingly, a growing number of studies have identified clinical associations between these nonca­
nonical SPs and disease, suggesting that there is functional significance to the alteration of SP backbone
structure. For example, associations have been found between SP chain length and cardiovascular disease, pain,
diabetes, and dementia. This review will provide an overview of the processes that are known to regulate
noncanonical SP accumulation, describe the clinical correlations reported for these molecules, and review the
experimental evidence for the potential functional implications of their dysregulation. It is likely that further
scrutiny of noncanonical SPs may provide new insight into pathophysiological processes, serve as useful bio­
markers for disease, and lead to the design of novel therapeutic strategies.

1. Introduction to noncanonical SPs several decades. However, one additional aspect of SP diversity that is
often overlooked involves variations of the structure of the sphingoid
Sphingolipids (SPs), including ceramides, sphingomyelins, ganglio­ backbone itself. In mammals, the majority of SPs contain a canonical
sides, cerebrosides, sulfatides, and others, represent a highly diverse 18‑carbon, dihydroxy, monounsaturated backbone, or “d18:1”. (See
class of lipids that are abundant in virtually all membranes of eukaryotic Fig. 1 for nomenclature.) As a result, most studies into SP functions
cells [1]. All SPs are built on an essential core structure; a sphingoid base disregard the quantitatively minor noncanonical species, and many
that serves as the scaffold for SPs of varying complexity [2]. The vast lipidomics studies do not discriminate canonical from noncanonical SPs.
diversity of SPs is due to the potential for the scaffold to be modified at This is particularly conspicuous regarding the widely studied signaling
two positions (Fig. 1A). Thousands of unique functional groups may be SP, sphingosine 1-phosphate (S1P). Despite the fact that “S1P” typically
added to the hydroxylated carbon at position 1, and dozens of struc­ refers to the aggregate mixture of several different S1P variants, func­
turally distinct acyl chains may be added to the amine group at carbon 2. tional studies almost uniformly address the canonical d18:1 without
This results in hundreds of empirically confirmed variants and tens of acknowledging potential roles of the quantitatively minor structural
thousands of additional species that are presumed or are necessary variants. For the purpose of this review, the term “noncanonical” refers
metabolic intermediates. This diversity is widely appreciated and has to any SPs that contain a sphingoid backbone other than d18:1 (Fig. 1B).
been scrutinized by mass spectrometry-based lipidomics studies for In Section 2, we first review the fundamentals of SP metabolism and the

* Corresponding author at: Department of Pharmacology, Yong Loo Lin School of Medicine, National University of Singapore, Singapore.
E-mail address: phcdrh@nus.edu.sg (D.R. Herr).

https://doi.org/10.1016/j.cellsig.2020.109890
Received 13 October 2020; Received in revised form 15 December 2020; Accepted 16 December 2020
Available online 28 December 2020
0898-6568/© 2020 Elsevier Inc. All rights reserved.
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

processes that provide structural diversity outside of the sphingoid 2.2. Generating diversity of canonical SPs
backbone. Section 3 describes the metabolic sources of noncanonical SPs
and emerging data that underscore their physiological and pathophys­ Dihydroceramides are formed by combining sphinganine with acyl-
iological importance. CoAs of varying chain lengths, which are then desaturated by dihy­
droceramide desaturase to form ceramides (Fig. 2). Ceramides can then
2. Sphingolipid metabolism act as a precursor to “higher-order” ceramides, such as glucosylcer­
amides, sphingomyelins and ceramide 1-phosphates, forming a meta­
2.1. De novo synthesis of the sphingoid backbone bolic “hub” with ceramides at the center [11].

The fundamental structural element of all sphingolipids is a sphin­ 2.2.1. Ceramides


goid base with an acyl chain that has some variation in length and As mentioned above, a major source of variation of canonical SPs
saturation. This makes up the sphingoid backbone which is formed by de occurs during the synthesis of ceramides, where acyl-CoAs are combined
novo synthesis in the endoplasmic reticulum (ER). Once produced in the with sphinganines, producing ceramides with different N-acyl chain
ER membrane, sphingolipids can be transported to the Golgi apparatus lengths (Fig. 1). This diversity is facilitated by the diversity of ceramide
for further packaging and processing, or to the plasma membrane where synthase (CerS) enzymes, 6 in total (CerS1-6), that vary in substrate
they participate in the maintenance of membrane architecture. preference [11,12]. Namely, Cer5/6 prefer C14-C18 acyl-CoAs, CerS1/4
The first and rate-limiting step of de novo synthesis is catalyzed by the prefer C18-C20 acyl-CoAs, and CerS2/3 prefer C18-C24 acyl-CoAs [11].
enzyme complex serine palmitoyltransferase (SPT), localized to the ER As a result, the human complement of ceramides contains N-linked acyl
membrane [3]. Initially characterized as a heterodimer [4], studies in chains that vary from 14 to 24 carbons, typically with 0 to 2 double
yeast have since demonstrated that SPT is made up of a heterotrimer of bonds, resulting in several dozen ceramide variants containing the same
main catalytic subunits, Lcb1p and Lcb2p, along with an additional canonical d18:1 SP backbone. Derangement from the “normal” com­
regulatory subunit, Tsc3p [5]. Human SPT is similarly made up of the plement of ceramide chain length has been associated with a number of
Lcb1p ortholog, SPTLC1, and two Lcb2p orthologs, SPTLC2 and SPTLC3 human diseases including cardiovascular disease, neurodegenerative
[6] (Fig. 2). Additionally, although there was no subunit found similar to diseases, diabetes, and cancer [11,13–15], suggesting that acyl chain
Tsc3p in mammals, two small short-membrane proteins were identified length is an important determinant of the pathological effects of cer­
to be functional orthologous, namely SPTssa and SPTssb [7]. These two amide [16,17].
proteins share the same function as Tsc3p, where their presence expo­ Ceramides can be further acylated at the 1-O-position or by
nentially increases the activity of the SPT complex [8]. ω-esterification to form acylceramides [18]. These ultra-long chain SPs
SPT is dependent on pyridoxal 5′ -phosphate (PLP) for activity, where are abundant in the skin and are known to be important for skin barrier
PLP is bound to the lysine on the SPTLC2/SPTLC3 subunits [4,9]. The function [19]. Due in large part to pioneering studies by Lina Obeid’s
lack of a PLP-binding motif in the SPTLC1 subunit [10] suggests that SPT group and others [20], modulation of acylceramides has been proposed
needs to exist as a complex of SPTLC1 with either SPTLC2 or SPTLC3 for as a mechanism of regulating the pro-apoptotic effects of ceramides.
SPT to carry out its function. Thus, SPT, as a heterotrimer, facilitates the Acylceramides are produced from ceramides on lipid droplets by the
condensation reaction between an amino acid and an acyl-CoA to form action of diacylglycerol acyltransferase 2 (DGAT2) to prevent an
3-ketosphinganines, which are rapidly reduced by ketosphinganine excessive accumulation of ceramides [20]. Thus, the generation of
reductase to “sphinganines,” representing the simplest SPs consisting acylceramides provides a protective effect against pro-apoptotic
only of a naked sphingoid backbone. As described below, this step is the ceramides which, when dysregulated, can give rise to a wide range of
sole determinant of the sphingoid backbone acyl chain length. On pathologies, such as non-alcoholic fatty liver disease [21] and obesity
average, the most common substrates for SPT are serine and the induced insulin resistance [22].
16‑carbon palmitoyl-CoA, catalyzing the formation of the canonical
sphingoid backbone and leading to the abundance of d18:1 2.2.2. Hexosylceramides
sphingolipids. Once formed by the acylation of the sphingoid backbone, ceramides

Fig. 1. Structure and variation of canonical and noncanonical mammalian SPs. A) This structure, showing the canonical d18:1 sphingoid backbone, indicates the
positions of the headgroup and N-acyl modifications that generate diversity among the canonical SPs. B) Using the canonical SP d18:1 sphingosine as a reference, this
structure shows where modifications occur within the sphingoid backbone to produce noncanonical variants. C1P, ceramide 1-phosphate; S1P, sphingosine
1-phosphate.

2
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

Fig. 2. Metabolic pathways producing noncanonical SPs in mammals. Variations in SP chain length and hydroxylation are due to differences in substrate use by the
SPT enzyme complex in the first step of de novo synthesis. The combination of SPT subunit combinations determines substrate preference. Resulting sphingoid
backbones enter the SP metabolic pathway and are converted to ceramides and higher-order ceramides. SDs are generated when FADS3 desaturates ceramides at
carbon position 14. SDs, sphingadienes; SP, sphingolipid; SPT, serine palmitoyltransferase.

then serve as scaffolds for the generation of a class of higher-order synthase 1 (SMS1) at the trans-Golgi lumen. Sphingomyelin can also be
ceramides known as the hexosylceramides. These are formed with the produced by sphingomyelin synthase 2 (SMS2) at the plasma membrane
attachment of a hexose sugar, glucose or galactose, to the primary hy­ [29], using ceramides produced locally at the membrane [30]. Since
droxyl group of ceramides by glucosylceramide synthase. These mono­ sphingomyelins are formed from ceramides by a single headgroup
hexosylceramides can be further modified with the additional hexose modification, the generation of sphingomyelins makes a relatively
sugars (glucose, galactose, mannose, fucose), other monosaccharides minor impact on overall SP diversity. However, the structural variation
(xylose), or amino sugars (sialic acid). The incorporation of up to a within the class of sphingomyelins is likely to be functionally mean­
dozen monosaccharides in straight-chain or branched configurations ingful [31,32].
results in the generation of hundreds of confirmed species [2] and
thousands of theoretical variants and metabolites. SP diversity increases
by orders of magnitude when hundreds to thousands of unique hexosyl 2.3. Degradation and recycling of the sphingoid backbone
headgroups are paired with the dozens of N-acyl ceramide variants.
Hexosylceramides can be further diversified by the addition of a sulfate Ceramides, either from de novo synthesis or from degradation of
to form sulfatides, known to be involved in neuronal differentiation and higher-order ceramides, can be readily hydrolyzed back into sphingo­
myelination [23]. Functionally, hexosylceramides localize in lipid rafts sine in what is known as the salvage pathway [33]. This pool of salvaged
where they participate in the regulation of cell signaling [24] and cell sphingosine can be re-acylated into ceramides and thus recycled back
differentiation [25]. into the hub of SP metabolism. This recycling contributes an estimated
50–90% of sphingolipid biosynthesis [34] and has been demonstrated to
2.2.3. Sphingomyelins be active in important cell signaling processes such as PKC signaling
Another class of higher-order SPs, the sphingomyelins, are formed [35]. Interestingly, it is known that ceramides produced by the salvage
from ceramide by the addition of a phosphocholine headgroup. Sphin­ pathway are functionally distinct from those generated by de novo syn­
gomyelins are the most abundant phospholipid present in most thesis [36]. Although the mechanism underlying this difference is un­
mammalian membranes, where they provide structural stability [26] clear, it is likely the result of differences in subcellular localization.
and contribute to the ordering of lipid rafts [27]. Ceramide is trans­ Sphingosines formed from the degradation of ceramides are phos­
ported from the ER lumen to the Golgi by a ceramide transfer protein, phorylated by two isoforms of sphingosine kinase (SPHK1 and SPHK2)
CERT [28], where sphingomyelin is synthesized by sphingomyelin to form sphingosine 1-phosphate (S1P). This process exists in dynamic
equilibrium, where S1P can be converted back to sphingosines by

3
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

selective S1P phosphatases (SGPP1 and SGPP2) [37–39] or by promis­ 3.1. Variations in sphingoid backbone chain length
cuous lipid phosphate phosphatases [40]. S1P can also be catabolized by
S1P lyase (SGPL1) to hexadecanal and ethanolamine 1-phosphate, Mammalian SPs largely contain backbones with a canonical
which feed into phospholipid biosynthesis [41]. It is important to note 18‑carbon, dihydroxy, monounsaturated acyl chain (d18:1), represent­
that hydrolysis of S1P by SGPL1 is the only known enzymatic process ing >60% of total SPs [49,50]. However, there is significant variation in
that irreversibly degrades the sphingoid backbone, thus making S1P a the backbone structures of quantitatively minor species [1]. The
key intermediate in SP degradation [42] and positioning SGPL1 as a following section will discuss variations in chain length, which in
critical regulator of homeostatic SP content [43]. In addition to this mammals, typically ranges from 14 to 22 carbons but species ranging
metabolic role, the biological roles of S1P have been extensively studied from 12 to 26 carbons have been reported [1]. Although most of these
since its identification as a potent cognate ligand of a family of G SPs have even-numbered chain lengths, odd-chain lengths are known to
protein-coupled receptors (S1P1 – S1P5). The effects of S1P receptor be present as well [50]. These odd-chain SPs will not be addressed in this
signaling are highly pleotropic and are reviewed in detail elsewhere review since their metabolic origins [51] and their structures (straight-
[44–46]. chain or branched) have not been confirmed [52] with one notable
exception, namely the 17‑carbon deoxymethylsphinganine, which is
3. Formation and biological roles of SPs with noncanonical discussed in Section 3.3.
sphingoid backbones
3.1.1. Metabolism of noncanonical chain lengths
The preceding sections describe mechanisms by which the canonical As mentioned above, SPT catalyzes the first step in de novo SP syn­
d18:1 sphingoid backbone is decorated to generate the structural di­ thesis with the generation of the basic building block for SPs: the
versity of SPs. However, there are several variations of the backbone sphingoid backbone. There are four typical configurations of the SPT
structure itself that have been known for decades, but whose signifi­ enzyme complex (Fig. 2), where each configuration has a different
cance has been largely neglected. This is due, in large part, to their preference for the length of acyl-CoA substrate [7]. The complex must
relatively low abundance and the technical challenges of quantification, contain SPTLC1 and one of two major non-obligatory subunits, SPTLC2
purification, and synthesis. However, due to improvements in analytical or SPTLC3 [4,6,10]. In addition, the incorporation of one of two small
techniques [47] and the formation of metabolomics consortia [48], subunits, SPTSSA (ssSPTa) and SPTSSB (ssSPTb), is necessary for full
emerging data suggest that these noncanonical SPs are precisely regu­ enzymatic activity [7]. In general, complexes containing SPTLC2 prefer
lated and that they play functional roles in physiology and disease a 16‑carbon substrate (palmitoyl-CoA) while SPTLC3 is more promis­
(Table 1). cuous [53,54], and SPTSSB prefers longer chain substrates, while
SPTSSA prefers shorter chains [7] (Fig. 2). The ability of the small
subunits to discriminate among acyl-CoA substrates of different chain
lengths is due to a single amino acid difference between SPTSSA and
Table 1
SPTSSB [8], where at the 25th amino acid position, SPTSSA has
Reported biological functions of noncanonical SPs.
methionine while SPTSSB has valine. Although the SPT complex is
Species Biological effect References present in all tissues, its three large subunits have differential but
Variations in sphingoid backbone chain length overlapping patterns of expression [6]. Notably, SPTLC3 expression is
d20:1 GM1 Neuronal differentiation and senescence [67–70] more restricted with highest expression in placental tissue [53]. This
suggests that 16‑carbon sphingolipids, made almost exclusively by
d16:1 S1P Inhibits cytokine expression in astrocytes [71]
SPTLC3 [54], require tissue-specific regulation but may be particularly
d20:1 S1P Inhibits COX-2 expression in glioblastoma [72] important during fetal development.
cells
Numerous studies have demonstrated that the length of the N-linked
Variations in sphingoid backbone saturation acyl chain of ceramides and sphingomyelins affects their biophysical
properties. This influences the rigidity/fluidity, permeability, and
d18:2 HexCer Inhibit viability of colon cancer and [89,100,101]
(plant- neuroblastoma cells through suppression of
thickness of resulting membranes, and alters membrane-protein in­
derived) AKT and WNT signaling teractions [55,56]. For example, membrane order is increased by the
incorporation of ceramides with longer N-acyl chains [57]. Interest­
d18:2 SPH [102]
(plant- ingly, similar positive correlations were observed between membrane
derived) order and the length of the ceramide sphingoid base [58]. In fact, it
d18:2 Cer Inhibits neurite outgrowth [105]
appears that the impact of sphingoid base length is more significant than
(plant- the length of the N-acyl chain with respect to biophysical properties of
derived) SPs [59].

Variations in sphingoid backbone hydroxylation 3.1.2. Known functions of SPs with noncanonical chain lengths
m18:0 SPH Disrupt mitochondrial dysfunction and [110,120–122]
m17:0 SPH promote neurodegeneration 3.1.2.1. Aging/neurodegeneration. Gangliosides, a class of higher-order
m18:1 Cer Neurotoxic via inhibition of SPHK and [124,125] hexosylceramides, localize within the outer layer of the plasma mem­
m18:1 SPH disruption of S1P receptor signaling brane and are mainly found in the central nervous system (CNS). They
m18:0 SPH Cytotoxic to pancreatic β islet cells and [129] confer structural rigidity to cells and form membrane microdomains
inhibits glucose-stimulated insulin secretion which mediate cell adhesion [60]. Additionally, they have many regu­
t20:0 Cer Promotes chemotaxis and viability in dermal [142–144] latory roles in the CNS, including the regulation of neurotrophic and
fibroblasts via JNK and Akt signaling neuroprotective properties pathways [61–64]. For example, the inter­
t18:0 S1P Promotes survival and maturation of oocytes [146] action of GM1, a member of the ganglioside family, with TrkA can
protect against neurotoxic stimuli and promote cell differentiation via
18:0 SPH Inhibits EMT and depletes cancer stem cells [147]
activation of the ERK1/2 pathway [62]. GM1 has also been shown to
18:0 SPH Inhibits melanin production in melanocytes [148,149] exert a protective effect against MPTP-induced oxidative stress and cell
Cer, ceramide; GM1, GM1 ganglioside; HexCer, hexosylceramide; S1P, sphin­ death in Neuro-2a neuroblastoma cells [63]. This could be due to the
gosine 1-phosphate, SPH, sphingosine.

4
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

role of GM1 as a mitochondrial regulator, whereby it promotes mito­ Cumulatively, these studies suggest that perturbations of SP metabolism,
chondriogenesis and enhances mitochondrial activity [65]. rather than total ceramide content, may be particularly relevant to the
Interestingly, both d18:1 and d20:1 GM1 are abundant in the CNS pathophysiology of CAD.
[66]. Currently, whilst there are no reports directly demonstrating
functional difference between the two naturally occurring GM1 species, 3.1.2.3. Insulin resistance. Several studies have drawn clinical and
correlation studies have nevertheless suggested functional relevance. functional correlations between ceramides and insulin resistance, a
For example, d18:1 and d20:1 GM1 are distributed non-homogenously major risk factor for T2DM (Table 2). For example, ceramide production
in the brain [67,68]. Although the absolute content of d18:1 GM1 is mediates TLR4-dependent induction of insulin resistance in vivo [86].
higher throughout the cortical regions, the relative enrichment of d20:1 Moreover, ceramides are shown be increased in a rat model of high-fat
GM1 is highest in the molecular layer of the dentate gyrus and lowest in diet-induced non-alcoholic fatty liver disease (NAFLD), an insulin
the amygdala, and the cerebral cortex has a gradient of d20:1 GM1 resistance-related hepatic disorder, and inhibition of ceramide synthesis
distribution, with the highest concentrations in the superficial layers reduces NAFLD-induced insulin resistance and hepatic lipid accumula­
[67,69]. Interestingly, in the aforementioned rat studies, d20:1 GM1 tion [87]. Similarly, decreasing the concentration of ceramides was
content increased markedly from postnatal day 0 to 3 months, and to a shown to improve glucose tolerance and insulin sensitivity in mice fed a
lesser extent from 3 months to 20 months, suggesting a role in both high fat diet [88]. Interestingly, a recent study demonstrated that the
neuronal differentiation and senescence. This is consistent with a pre­
vious study reporting that d20:1 gangliosides were detectable in adult
brain, but were undetectable in embryonic brain tissues from multiple Table 2
species including human [70]. Reported clinical correlations between noncanonical SPs and disease or de­
To determine whether specific, noncanonical S1P species correlate mographic characteristics.
with neurodegenerative disease, we evaluated the four most abundant Species Clinical correlation References
S1P species in plasma from patients with Alzheimer’s disease (AD) or Variations in sphingoid backbone chain length
vascular dementia (VaD) relative to cognitively normal controls [71].
d16:1 S1P ↓ Vascular dementia [71]
We found that specific decreases in d16:1 S1P were associated with VaD,
while no associations were found with AD. We further demonstrated d16:1 S1P ↑ Oxaliplatin treatment [74]

that d16:1 S1P correlated negatively with inflammatory cytokines while d20:0 SPH ↑ Cardiovascular event [83]
canonical d18:1 S1P correlated positively. Surprisingly, in vitro studies d16:1/16:0 Cer ↑ HOMA-IR [15]
demonstrated that both d16:1 and d18:1 S1P induce cytokine expression d16:1/18:0 Cer
in astrocytic cells. However, the efficacy of d16:1 S1P was lower, and d16:1/20:0 Cer
was able to attenuate the inflammatory effect of d18:1 S1P when both d16:1/18:0 SM
d16:1/20:0 SM
species were administered simultaneously. It remains unclear whether d16:1 SPH
this phenomenon occurs in vivo, since endogenous levels of d16:1 S1P d16:1 S1P
are low to undetectable in the brain [49]. However, another group re­ d17:1 S1P
ported a similar finding for d20:1 S1P, which is present in significant d16:1/18:0 SM ↑ Type 2 Diabetes Mellitus [15]
quantities in the brain [72]. That study demonstrates that d18:1 S1P-
mediated induction of inflammatory COX-2 expression in glioblastoma Variations in sphingoid backbone saturation
cells can be attenuated by co-administration of d20:1 S1P, also sup­ d18:2/24:1 Cer ↑ Aging [152]
porting a model that noncanonical S1Ps may play a role in “fine tuning” d18:2/16:0 SM
S1P-mediated inflammation. d18:2/23:0 Cer ↓ HOMA-IR [15,152]
The different inflammomodulatory effects of individual S1P species d18:2/24:0 Cer
are likely to be the result of differential interactions with their cognate d18:2/24:1 Cer
receptors. Indeed, although all S1P species tested have been shown to d18:2/25:0 Cer
d18:2/26:0 Cer
interact with cognate S1P receptors, they display differences in their
d18:2/18:1 SM
potencies and efficacies [73]. For example, shortening the S1P acyl d18:2/23:0 SM
chain (<18 carbons) increases its efficacy toward S1P4, but decreases its d18:2/24:0 SM
efficacy toward all other cognate receptors. Our previous study supports d18:2/16:0 Hex2Cer
this model by demonstrating that patients receiving cisplatin exhibited a d18:2/22:0 HexCer
d18:2/23:0 HexCer
specific increase in plasma d16:1 S1P content, but not other S1P species d18:2/24:0 HexCer
[74]. We further show that activation of receptor S1P2 attenuates d18:2/24:1 HexCer
cisplatin-mediated neuropathy and that d16:1 S1P is selectively less
d18:2/18:0 SM ↑ HOMA-IR [15]
potent toward S1P2. We hypothesize that elevation of d16:1 S1P pref­ d18:2 SPH
erentially activates pro-neuropathic S1P1 [75–78] at the expense of d18:2 S1P
neuroprotective S1P2. m18:2 SPH ↓ Cardiovascular Event [83]

3.1.2.2. Cardiovascular disease. SPs have been shown to act as car­ Variations in sphingoid backbone hydroxylation
dioprotectants at physiological concentrations, but an increase in SP m18:0 SPH ↑ HSN Type I (A and C) [111]
content, particularly ceramides, can be toxic to many cell types m17:0 SPH
including cardiomyocytes [79–82]. Interestingly, a study published in m18:1 Cer ↑ CIPN [123]
2015 has shown that only elevated d20:1 sphingosine, but not canonical m18:1 SPH
d18:1 sphingosine, was associated with an increased risk of coronary m18:0 SPH ↑ Type II Diabetes Mellitus [107,127,128]
artery disease (CAD) [83]. In addition, another recent study identified
m18:1 SPH
positive correlations between d16:1/18:1 ceramide and both large artery
atherosclerosis and cerebrovascular disease [84]. Due to the relative high t20:0 Cer ↓ Psoriasis [137,153]
expression of SPTLC3 in myocardium [85], it is likely that these non­ ↓ Atopic dermatitis

canonical SPs are produced by pathological conditions in the heart itself. Cer, ceramide; HexCer, hexosylceramide; Hex2Cer, dihexosylceramide; S1P,
sphingosine 1-phosphate; SM, sphingomyelin; SPH, sphingosine.

5
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

protective effect of sulforaphane against insulin resistance is likely to be their d18:1 counterparts. Interestingly, women demonstrated signifi­
the result of inhibition of SPTLC3 [88]. Although this study demon­ cantly higher concentrations of SDs compared to men [15], suggesting
strated only that SPTLC3 inhibition reduced total ceramide content and that a protective effect of SDs may contribute to the lower incidence of
did not evaluate individual species, it is likely that noncanonical CAD in women.
ceramides were preferentially depleted due to SPTLC3’s broad substrate In addition to the clinical correlations, a number of preclinical
specificity. This is consistent with a large scale clinical study conducted studies demonstrated that SDs have significant biological effects on
by our group that demonstrated that d16:1 ceramide species, have mammalian cells. For example, soy-based 4E,8E-d18:2 hexosylceramide
stronger positive correlations with both insulin resistance and body was shown to inhibit colon tumorigenesis and reduce gene expression of
mass index compared to canonical d18:1 ceramides [15]. cancer-related transcription factors in two different murine cancer
models [100]. Similarly, 4E,8Z-d18:2 sphingosine (soy) and 4E,6E-d18:2
3.2. Variations in sphingoid backbone saturation sphingosine (insect) were shown to reduce the viability of colon cancer
cells in vitro and in vivo [89] and can downregulate tumorigenic Wnt
The canonical d18:1 sphingoid backbone contains a trans double signaling [101]. Similar antiproliferative effects of soy-based SDs were
bond between carbons 4 and 5 (4E-d18:1 or d18:1∆4t), but naturally demonstrated in cultured neuroblastoma cells and in a neuroblastoma
occurring species contain 0 to 2 double bonds. Fully saturated SPs xenograft model [102]. Coupled with the demonstration that soy-based
(d18:0) are common and are, in fact, a necessary intermediate in the SDs can be absorbed though the gut and enter peripheral circulation
production of d18:1 species (Fig. 2). Sphingoid backbones containing [103,104], these data suggest that dietary plant-based SDs may be
two double bonds (d18:2), the sphingadienes (SDs), are present in many broadly chemopreventive.
mammalian tissues but at a relatively low abundance. Interestingly, the SDs have also been shown to exert effects on neural cells. For
existence of SDs has been shown in many species, but the position of the example, d18:2 ceramides derived from the edible tuber konjac can
second double bond demonstrates characteristic phylogenetic differ­ inhibit nerve growth factor-mediated neurite outgrowth via an Sema3A-
ences. Plants, such as soy, are characterized by an abundance of ∆4,∆8 dependent mechanism [105]. Interestingly, konjac contains both 4E,8Z-
SDs [89] while insects contain an unusual conjugated ∆4,∆6 configu­ d18:2 and 4E,8E-d18:2 ceramides. While both confirmations can inhibit
ration [90]. To date, the only specific SD species reported in humans neurite outgrowth, only 4E,8E-d18:2 ceramide can bind Sema3A,
contain a 4E,14Z-d18:2 backbone [1]. providing evidence for mechanistically distinct functions of different
Fatty acid saturation has a marked effect on the behavior of SPs in a noncanonical backbones.
fluid membrane, which consequently impacts membrane properties. It
has been previously established that introduction of cis double bonds 3.3. Variations in sphingoid backbone hydroxylation
into the N-linked acyl chain of ceramide reduces membrane order and
rigidity [91]. More recently, a similar effect was seen with the sphingoid The third known backbone modification that results in noncanonical
backbone in that 4E,14Z-d18:2 ceramides destabilized membrane gel SPs involves the number of hydroxyl moieties. Canonical dihydroxy
phases relative to that of canonical d18:1 ceramides [92]. This raises the d18:1 SPs contain two hydroxyl groups at carbon positions 1 and 3, but
possibility that accumulation of SP dienes results in the dispersion of SP- monohydroxy (m18:1) and trihydroxy (t18:0) species are known to exist
rich membrane microdomains, potentially disrupting signaling protein in mammals. The most common of these noncanonical variants is
complexes [93]. “phytosphingosine” (t18:0), which, although detectable in mammals, is
much more abundant in plants and fungi.
3.2.1. Metabolism of sphingadienes
Although the existence of d18:2 SPs in humans has been known since 3.3.1. Metabolism of hydroxylation variants of the sphingoid backbone
the 1960’s [94,95], the metabolic source of this backbone structure Although SPT largely uses the amino acid serine as a preferred sub­
remained elusive until 2019. Two independent groups reported that the strate for condensation with acyl-CoA, it is also able to utilize other
enzyme fatty acid desaturase 3 (FADS3) can catalyze the formation of amino acids such as alanine and glycine (Fig. 3), resulting in the rela­
4E,14Z-d18:2 SDs [96,97]. Although the precise substrate preference tively inefficient formation of deoxy-SPs [106]. Condensation of L-
has not been described in detail, it is likely that FADS3 desaturates d18:1 alanine with palmitoyl-CoA forms deoxysphinganine (m18:0) and the
ceramides rather than sphingosines. The resulting d18:2 ceramides enter condensation of glycine with palmitoyl-CoA forms deoxymethyl­
the SP metabolic pathway leading to the accumulation of higher-order sphinganine (m17:0) [107]. Deoxysphinganines and deoxymethyl­
d18:2 ceramides. d18:2 sphingosine can be efficiently phosphorylated sphinganines then serve as substrates for ceramide synthase to form
by both SPHK1 and SPHK2 to form d18:2 S1P [98,99]. However, d18:2 deoxyceramides and deoxymethylceramides, respectively [108]. It is
S1P is less susceptible to terminal degradation by SGPL1 relative to suggested that the utilization of alternative amino acids occurs due to a
d18:1 [97], suggesting increased metabolic stability of SDs. serine deficiency in the extracellular environment [109,110], leading to
elevated levels of deoxysphingolipids [111,112]. In addition, poly­
3.2.2. Known functions of SPs with sphingadiene backbones morphisms within the SPTLC1 and SPTLC2 subunits of SPT have been
Although the functional roles of SDs in mammalian physiology/ shown to result in substrate promiscuity leading to increased production
pathophysiology are not yet clear, their asymmetric tissue distribution of deoxysphingolipids [113–116]. Although it is not clear whether the
suggests the potential significance of maintaining SD homeostasis. small subunits of SPT participate in specificity of the amino acid sub­
Notably, an analysis of mouse tissues demonstrated that SD ceramides strate, it has been reported that, relative to SPTSSA, SPTSSB is a more
are particularly abundant in the kidney and brain, where they represent efficient activator of a mutant form of SPTLC1 (SPTLC1S331F) that
68% and 30% of total ceramides, respectively [97]. Furthermore, clin­ demonstrates substrate promiscuity [8]. Furthermore, a yeast study
ical studies implicate their potential role in the pathophysiology of demonstrated that the functional homolog of SPT small subunits (Tsc3),
cardiovascular and metabolic disease (Table 2). The study described in addition to activating SPT, also plays an important role in amino acid
above that identified a positive relationship between d20:1 sphingosine selectivity [114] allowing yeast SPT to use alanine as a substrate.
and CAD also identified a negative association between d18:2 sphingo­ Interestingly, instead of the trans- 4E double bond present in ca­
sine and subsequent cardiovascular event (HR = 0.79) [83]. This sug­ nonical sphingolipids, 1-deoxysphingosines possess a cis- 14Z double
gests that SDs may play a protective role in cardiovascular disease. In bond [117], potentially affecting the metabolism and functional activity
addition, our large-scale lipidomics analysis [15] demonstrated that of m18:1 SPs relative to that of d18:1 species. As described above, 14Z is
d18:2 SD species correlate negatively with obesity and insulin resis­ the position and confirmation of the second double bond in the more
tance, or in some cases, showed weaker positive correlations relative to abundant d18:2 SD species (see Section 3.2), and it was confirmed that

6
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

Fig. 3. Metabolic pathways producing deoxy-SPs in mammals. The generation of deoxy-SPs is typically rare in mammals, but certain mutations of SPT result in
increased preference for alanine or glycine substrates, resulting in the absence of hydroxylation at carbon position 1. HSAN1, hereditary sensory and autonomic
neuropathy type 1.

this desaturation reaction is catalyzed by the SD-producing enzyme with L-serine, thus favoring production of canonical d18:1 SPs [109].
FADS3 [96]. The absence of 4E-m18:1 SPs suggests that m18:0 SPs Deoxy-SPs may be involved in other, non-hereditary forms of neu­
cannot serve as substrates for dihydroceramide desaturase. Due to the ropathy. For example, an association between deoxy-SPs and CIPN was
absence of the C1 hydroxyl group, m18:1 SPs also cannot serve as sub­ identified in a cohort of 27 breast cancer patients treated with paclitaxel,
strates for glucosylceramide synthases or sphingomyelin synthases and where it was found that several species of m18:1 ceramides correlated
thus cannot be metabolized into higher-order ceramides [107]. Simi­ positively with the incidence and severity of neuropathy [123]. Mech­
larly, they cannot serve as substrates for sphingosine kinases to form anistically, it was demonstrated that m18:1 ceramides and m18:1
S1Ps for their subsequent degradation by S1P lyase [106]. As a result, sphingosine were elevated in the dorsal root ganglia (DRGs) of mice
these deoxysphingolipids are metabolically stable and accumulate in treated with docetaxel in a model for chemotherapy-induced peripheral
tissues as ceramides or free sphingoid bases, lending to their neurotox­ neuropathy (CIPN) [124]. The resulting m18:1 sphingosine-induced
icity (see below). neurotoxicity (neurite swelling) could be ameliorated by co-
The metabolic origin of t18:0 phytosphingosine in mammals is less administration of canonical d18:1 S1P, suggesting that the mechanism
well characterized and may be obtained primarily through dietary underlying m18:1 SP-induced neurotoxicity may, in part, be the result of
sources [118]. However, is has been demonstrated that ∆4-sphingolipid the disruption of S1P receptor signaling. Interestingly, it was shown that
desaturate (DEGS2), the enzyme responsible for introducing the 4E derivatives of deoxy-sphinganine (m18:0 sphingosine) selectively
double bond, is also able to instead hydroxylate the same position [119], inhibit SPHK1 [125,126], offering a potential pharmacophore for the
resulting in t18:0–1,3,4-triol phytosphingosine. The relative contribu­ development of SPHK1-targeted drugs. It is likely, therefore, that
tion of DEGS2 to the accumulation of t18:0 SPs in mammals in unclear. disruption of S1P signaling due to accumulation of deoxy-SPs is the
result of the direct inhibition and proteasomal degradation of SPHK1,
3.3.2. Known functions of deoxy-SPs leading to local and intracellular depletion of S1P.
The clearest evidence that deoxy-SPs have functional relevance in Elevated plasma deoxy-SPs are also found to be associated with pa­
human pathophysiology is their known associations with disease, most tients with type 2 diabetes mellitus (T2DM) but not with T1DM
notably, hereditary sensory and autonomic neuropathy type I (HSAN I). [127,128]. m18:0 and m18:1 sphingosine improved the predictive value
Certain forms of this hereditary syndrome are associated with missense of blood glucose for incident diabetes, but their predictive power
mutations of SPTLC1 or SPTLC2 resulting in the accumulation of deoxy- declined with increasing BMI and was only significant for non-obese
SPs [120]. Persistent elevation of these noncanonical SP species results individuals with BMI <30 [127]. It is likely that accumulation of
in neurotoxicity, resulting in Ca2+ mishandling and neurodegeneration deoxy-SPs contributes mechanistically to the progression of T2DM since
[121], likely through disruption of mitochondrial function m18:0 sphingosine has been shown to be cytotoxic to pancreatic β islet
[111,120,122]. In HSN1 patients, accumulation of deoxy-SPs and cells and inhibited glucose-stimulated insulin secretion [129]. It is
resultant disease severity can be attenuated by dietary supplementation notable, however, that HSAN1 patients, who present with chronic

7
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

elevation of m18:0 sphingosine, are not characterized by overt T2DM mesenchymal transition in breast cancer cells and reduce cancer stem
[130]. This suggests that, although deoxy-SPs probably contribute to cell proliferation in a spheroid assay, likely by binding and antagonizing
T2DM progression, they are not sufficient for disease onset. Considering EGF receptor [147]. Interestingly, t18:0 phyto-sphingosine was also
that deoxy-SPs are positively associated with T2DM and are neurotoxic, shown to inhibit melanin production in human skin cells, presumably
it is tempting to speculate whether they play a functional role in the though the inhibition of microphthalmia-associated transcription factor
development of diabetic neuropathy. However, a study of 39 patients [148]. Although this study was performed in the context of cosmetic
with diabetic distal sensorimotor polyneuropathy showed an expected applications, it suggests that t18:0 phyto-sphingosine may potentially be
increase in plasma m18:1 sphingosine relative to healthy non-diabetic protective against melanoma [149].
controls, but this study showed no correlation between deoxy-SP con­ Although there are no reports implicating t18:0 phyto-SPs as direct
tent and clinical severity of neuropathy [131]. regulators of neurophysiological responses, it is notable that one recent
Although the causal relationships between deoxy-SPs and diabetes or study found that plasma t18:0 phyto-sphingosine was identified as a
non-hereditary neuropathy are not known, it is worth noting that potential biomarker for clinical depression using a bioinformatics
metabolic perturbance can favor deoxy-SP production. Specifically, approach [150]. While this does not define a causal relationship, it is
experimental serine deficiency was shown to induce the elevation of intriguing to speculate whether this may involve gut-brain axis in­
deoxy-SPs in vitro and in vivo [110], and in humans, plasma deoxy-SP teractions [151].
content correlates negatively with serine concentration [112]. Inter­
estingly, in some cancers, a source of exogenous serine is required to 4. Forward looking statement
prevent the endogenous production of cytotoxic deoxy-SPs [132]. One
source of exogenous serine may be sensory neurons innervating the Although the marked structural diversity of SPs has been understood
tumor [133]. It is tempting to speculate, therefore, that depletion of since the mid-20th century, the functional significance of the quantita­
serine by highly metabolic cancer cells may result in local conditions tively minor noncanonical species has only been addressed within the
that favor neuronal deoxy-SP production, contributing to cancer-related last 20 years. Variation of the sphingoid backbone will necessarily affect
neuropathy. the biophysical properties of membrane bilayers, but in addition,
emerging data demonstrate that the diversity of backbone structure also
3.3.3. Known functions of t18:0 “phyto” SPs influences cell signaling. Perhaps the clearest example of this is
Phyto-SPs are particularly abundant in the skin and are likely to play demonstrated by the differences in receptor selectivities among the S1Ps
significant roles in barrier integrity and skin homeostasis [134]. For [72–74,141], and the fact that these differences affect meaningful
example, treatment of human keratinocytes with t18:0 phytosphingo­ changes in cell behavior [71,72,146]. Moving forward, it is likely that
sine has been shown to cause an increase in t18:0 phytoceramides [135]. increasingly sensitive mass spectrometry technologies will allow routine
This is likely to be functionally relevant since it has been shown that evaluation of noncanonical SPs, resulting in the identification of new
t18:0 phytoceramides are significantly lower in skin lesions associated clinical correlations, and thus, leading to the identification of novel
with psoriasis [136] and atopic dermatitis [137], and that increasing the biological functions. We, the authors, hope that this review helps to
skin content of t18:0 phytoceramides improves the symptoms of atopic draw attention to the potential roles of noncanonical SPs and that future
dermatitis [138]. This is likely to be mediated by the production of studies will place increased emphases on these often-overlooked mo­
natural moisturizing factor, possibly due to the activation of the lecular mediators.
filament-associated protein filaggrin [139].
Accumulation of phytoceramides will result in the production of Acknowledgements
t18:0 phyto-S1P, therefore, it is likely that many of the effects of phyto-
SPs on the skin are mediated by this signaling molecule. Although its This work was supported by grants from the Ministry of Education,
ability to activate the entire S1P receptor family is assumed, it remains Singapore (R-181-000-183-114, WYO and DRH), the National Univer­
incompletely characterized. t18:0 phyto-S1P has been shown to bind at sity Health System, Singapore (NUHSRO/2019/051/T1/Seed-Mar/04,
least to both S1P1 and S1P4 [140], and can activate S1P4 with ~50-fold WYO and DRH), and the National Medical Research Council, Singapore
greater potency compared to S1P [141]. t18:0 phyto-S1P promoted (NMRC/CSA-SI/007/2016, CPC, MKPL, and DRH).
chemotaxis in murine dermal fibroblasts in pertussis toxin-sensitive
manner, presumably via the activation of either S1P2 or S1P4 [142]. In References
human dermal fibroblasts, t18:0 phyto-S1P was shown to increase cell
viability through the regulation of JNK and AKT signaling [143]. Since [1] S.T. Pruett, A. Bushnev, K. Hagedorn, M. Adiga, C.A. Haynes, M.C. Sullards, D.
C. Liotta, A.H. Merrill, Thematic review series: sphingolipids. Biodiversity of
these cells express S1P1, S1P2, and S1P3, but not S1P4, the effects of sphingoid bases (“sphingosines”) and related amino alcohols, J. Lipid Res. 49
phyto-t18:0 S1P are presumably mediated via the activation of one or (2008) 1621–1639, https://doi.org/10.1194/jlr.R800012-JLR200.
more of these S1P receptors. Cumulatively, these two studies suggest the [2] E. Fahy, S. Subramaniam, H.A. Brown, C.K. Glass, A.H. Merrill, R.C. Murphy, C.R.
H. Raetz, D.W. Russell, Y. Seyama, W. Shaw, T. Shimizu, F. Spener, G. van Meer,
potential importance of S1P2 in t18:0 phyto-S1P signaling, although this M.S. VanNieuwenhze, S.H. White, J.L. Witztum, E.A. Dennis, A comprehensive
has yet to be confirmed. Interestingly, phyto-t18:0 S1P synergizes classification system for lipids, J. Lipid Res. 46 (2005) 839–862, https://doi.org/
epidermal growth factor (EGF) signaling, and their combined topical use 10.1194/jlr.E400004-JLR200.
[3] S. Pyne, N.J. Pyne, Sphingosine 1-phosphate signalling in mammalian cells,
has been shown to improve moisture content, dermal thickness, and skin Biochem. J. 349 (2000) 385, https://doi.org/10.1042/0264-6021:3490385.
elasticity in a human clinical study [144]. [4] K. Hanada, Serine palmitoyltransferase, a key enzyme of sphingolipid
Since these noncanonical SPs are abundant in certain bacteria and metabolism, Biochim. Biophys. Acta BBA - Mol. Cell Biol. Lipids 1632 (2003)
16–30, https://doi.org/10.1016/S1388-1981(03)00059-3.
fungi, and appear to be produced only in small quantities by mammalian
[5] K. Gable, H. Slife, D. Bacikova, E. Monaghan, T.M. Dunn, Tsc3p is an 80-amino
cells, it is likely that the skin microbiome is the primary source of acid protein associated with serine palmitoyltransferase and required for optimal
bioactive t18:0 SPs, and that these “postbiotics” are relevant mediators enzyme activity, J. Biol. Chem. 275 (2000) 7597–7603, https://doi.org/10.1074/
of the activity of commensal skin microorganisms [145]. jbc.275.11.7597.
[6] T. Hornemann, S. Richard, M.F. Rütti, Y. Wei, A. von Eckardstein, Cloning and
Phyto-SPs have been shown to have roles in tissues other than the initial characterization of a new subunit for mammalian serine-
skin. For example, t18:0 phyto-S1P promoted survival and maturation of palmitoyltransferase, J. Biol. Chem. 281 (2006) 37275–37281, https://doi.org/
porcine oocytes in vitro, likely though the SOD3-mediated attenuation of 10.1074/jbc.M608066200.
[7] G. Han, S.D. Gupta, K. Gable, S. Niranjanakumari, P. Moitra, F. Eichler, R.
oxidative stress, and the activation of ERK and AKT survival pathways H. Brown, J.M. Harmon, T.M. Dunn, Identification of small subunits of
[146]. Furthermore, there is evidence that phyto-SPs may be chemo­ mammalian serine palmitoyltransferase that confer distinct acyl-CoA substrate
protective. t18:0 phyto-sphingosine was shown to inhibit epithelial-

8
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

specificities, Proc. Natl. Acad. Sci. 106 (2009) 8186–8191, https://doi.org/ [33] M.K.P. Lai, W.S. Chew, F. Torta, A. Rao, G.L. Harris, J. Chun, D.R. Herr, Biological
10.1073/pnas.0811269106. effects of naturally occurring sphingolipids, uncommon variants, and their
[8] J.M. Harmon, D. Bacikova, K. Gable, S.D. Gupta, G. Han, N. Sengupta, analogs, NeuroMolecular Med. 18 (2016) 396–414, https://doi.org/10.1007/
N. Somashekarappa, T.M. Dunn, Topological and functional characterization of s12017-016-8424-8.
the ssSPTs, small activating subunits of serine palmitoyltransferase, J. Biol. [34] K. Kitatani, J. Idkowiak-Baldys, Y.A. Hannun, The sphingolipid salvage pathway
Chem. 288 (2013) 10144–10153, https://doi.org/10.1074/jbc.M113.451526. in ceramide metabolism and signaling, Cell. Signal. 20 (2008) 1010–1018,
[9] S. Yasuda, M. Nishijima, K. Hanada, Localization, topology, and function of the https://doi.org/10.1016/j.cellsig.2007.12.006.
LCB1 subunit of serine PALMITOYLTRANSFERASE in mammalian cells, J. Biol. [35] K.P. Becker, K. Kitatani, J. Idkowiak-Baldys, J. Bielawski, Y.A. Hannun, Selective
Chem. 278 (2003) 4176–4183, https://doi.org/10.1074/jbc.M209602200. inhibition of juxtanuclear translocation of protein kinase C βII by a negative
[10] K. Hanada, T. Hara, M. Nishijima, O. Kuge, R.C. Dickson, M.M. Nagiec, feedback mechanism involving ceramide formed from the salvage pathway,
A mammalian homolog of the yeast LCB1 encodes a component of serine J. Biol. Chem. 280 (2005) 2606–2612, https://doi.org/10.1074/jbc.
palmitoyltransferase, the enzyme Catalyzing the first step in sphingolipid M409066200.
synthesis, J. Biol. Chem. 272 (1997) 32108–32114, https://doi.org/10.1074/ [36] S. Hebbar, I. Sahoo, A. Matysik, I. Argudo Garcia, K.A. Osborne, C. Papan,
jbc.272.51.32108. F. Torta, P. Narayanaswamy, X.H. Fun, M.R. Wenk, A. Shevchenko, D. Schwudke,
[11] T.D. Mullen, Y.A. Hannun, L.M. Obeid, Ceramide synthases at the Centre of R. Kraut, Ceramides and stress signalling intersect with autophagic defects in
sphingolipid metabolism and biology, Biochem. J. 441 (2012) 789–802, https:// neurodegenerative drosophila blue cheese (bchs) mutants, Sci. Rep. 5 (2015)
doi.org/10.1042/BJ20111626. 15926, https://doi.org/10.1038/srep15926.
[12] Y.A. Hannun, L.M. Obeid, Many ceramides, J. Biol. Chem. 286 (2011) [37] K.R. Johnson, K.Y. Johnson, K.P. Becker, J. Bielawski, C. Mao, L.M. Obeid, Role of
27855–27862, https://doi.org/10.1074/jbc.R111.254359. human sphingosine-1-phosphate phosphatase 1 in the regulation of intra- and
[13] M. Ding, K.M. Rexrode, A review of Lipidomics of cardiovascular disease extracellular sphingosine-1-phosphate levels and cell viability, J. Biol. Chem. 278
highlights the importance of isolating lipoproteins, Metabolites 10 (2020) 163, (2003) 34541–34547, https://doi.org/10.1074/jbc.M301741200.
https://doi.org/10.3390/metabo10040163. [38] C. Ogawa, A. Kihara, M. Gokoh, Y. Igarashi, Identification and characterization of
[14] P.L. Wood, Lipidomics of Alzheimer’s disease: current status, Alzheimers Res. a novel human sphingosine-1-phosphate phosphohydrolase, hSPP2, J. Biol.
Ther. 4 (2012) 5, https://doi.org/10.1186/alzrt103. Chem. 278 (2003) 1268–1272, https://doi.org/10.1074/jbc.M209514200.
[15] W.S. Chew, F. Torta, S. Ji, H. Choi, H. Begum, X. Sim, C.M. Khoo, E.Y.H. Khoo, [39] C. Mao, J.D. Saba, L.M. Obeid, The dihydrosphingosine-1-phosphate
W.-Y. Ong, R.M. Van Dam, M.R. Wenk, E.S. Tai, D.R. Herr, Large-scale lipidomics phosphatases of Saccharomyces cerevisiae are important regulators of cell
identifies associations between plasma sphingolipids and T2DM incidence, JCI proliferation and heat stress responses, Biochem. J. 342 (1999) 667–675, https://
Insight. 5 (2019), https://doi.org/10.1172/jci.insight.126925. doi.org/10.1042/bj3420667.
[16] S. Grösch, S. Schiffmann, G. Geisslinger, Chain length-specific properties of [40] D. Saba Julie, Hla Timothy, Point-counterpoint of sphingosine 1-phosphate
ceramides, Prog. Lipid Res. 51 (2012) 50–62, https://doi.org/10.1016/j. metabolism, Circ. Res. 94 (2004) 724–734, https://doi.org/10.1161/01.
plipres.2011.11.001. RES.0000122383.60368.24.
[17] J.L. Stith, F.N. Velazquez, L.M. Obeid, Advances in determining signaling [41] P. Bandhuvula, J.D. Saba, Sphingosine-1-phosphate lyase in immunity and
mechanisms of ceramide and role in disease, J. Lipid Res. 60 (2019) 913–918, cancer: silencing the siren, Trends Mol. Med. 13 (2007) 210–217, https://doi.
https://doi.org/10.1194/jlr.S092874. org/10.1016/j.molmed.2007.03.005.
[18] M. Rabionet, A. Bayerle, C. Marsching, R. Jennemann, H.-J. Gröne, Y. Yildiz, [42] W. Stoffel, G. Assmann, Metabolism of sphingosine bases, XV. Enzymatic
D. Wachten, W. Shaw, J.A. Shayman, R. Sandhoff, 1-O-acylceramides are natural degradation of 4t-Sphingenine 1-phosphate (sphingosine 1-phosphate) to 2t-
components of human and mouse epidermis, J. Lipid Res. 54 (2013) 3312–3321, Hexadecen-1-al and ethanolamine phosphate, Biol. Chem. 351 (1970)
https://doi.org/10.1194/jlr.M040097. 1041–1049, https://doi.org/10.1515/bchm2.1970.351.2.1041.
[19] M. Akiyama, Acylceramide is a key player in skin barrier function: insight into the [43] A. Kumar, J. Zamora-Pineda, E. Degagné, J.D. Saba, S1P Lyase regulation of
molecular mechanisms of skin barrier formation and ichthyosis pathogenesis, thymic egress and oncogenic inflammatory signaling, Mediat. Inflamm. 2017
FEBS J. (2020), https://doi.org/10.1111/febs.15497 n/a (n.d.). (2017), e7685142, https://doi.org/10.1155/2017/7685142.
[20] C.E. Senkal, M.F. Salama, A.J. Snider, J.J. Allopenna, N.A. Rana, A. Koller, Y. [44] N.J. Pyne, S. Pyne, Recent advances in the role of sphingosine 1-phosphate in
A. Hannun, L.M. Obeid, Is Ceramide, Metabolized to Acylceramide and stored in cancer, FEBS Lett. (2020), https://doi.org/10.1002/1873-3468.13933.
lipid droplets, Cell Metab. 25 (2017) 686–697, https://doi.org/10.1016/j. [45] N.J. Pyne, S. Pyne, Sphingosine 1-phosphate receptor 1 signaling in mammalian
cmet.2017.02.010. cells, Mol. Basel Switz. 22 (2017), https://doi.org/10.3390/molecules22030344.
[21] J. Simon, A. Ouro, L. Ala-Ibanibo, N. Presa, T.C. Delgado, M.L. Martínez-Chantar, [46] M. Adada, D. Canals, Y.A. Hannun, L.M. Obeid, Sphingosine-1-phosphate receptor
Sphingolipids in non-alcoholic fatty liver disease and hepatocellular carcinoma: 2, FEBS J. 280 (2013) 6354–6366, https://doi.org/10.1111/febs.12446.
ceramide turnover, Int. J. Mol. Sci. 21 (2019) 40, https://doi.org/10.3390/ [47] B. Burla, S. Muralidharan, M.R. Wenk, F. Torta, Sphingolipid analysis in clinical
ijms21010040. research, Methods Mol. Biol. Clifton NJ. 1730 (2018) 135–162, https://doi.org/
[22] W.L. Holland, J.T. Brozinick, L.-P. Wang, E.D. Hawkins, K.M. Sargent, Y. Liu, 10.1007/978-1-4939-7592-1_11.
K. Narra, K.L. Hoehn, T.A. Knotts, A. Siesky, D.H. Nelson, S.K. Karathanasis, G. [48] B. Yu, K.A. Zanetti, M. Temprosa, D. Albanes, N. Appel, C.B. Barrera, Y. Ben-
K. Fontenot, M.J. Birnbaum, S.A. Summers, Inhibition of ceramide synthesis Shlomo, E. Boerwinkle, J.P. Casas, C. Clish, C. Dale, A. Dehghan, A. Derkach, A.
ameliorates glucocorticoid-, saturated-fat-, and obesity-induced insulin H. Eliassen, P. Elliott, E. Fahy, C. Gieger, M.J. Gunter, S. Harada, T. Harris, D.
resistance, Cell Metab. 5 (2007) 167–179, https://doi.org/10.1016/j. R. Herr, D. Herrington, J.N. Hirschhorn, E. Hoover, A.W. Hsing, M. Johansson, R.
cmet.2007.01.002. S. Kelly, C.M. Khoo, M. Kivimäki, B.S. Kristal, C. Langenberg, J. Lasky-Su, D.
[23] A.H. Merrill, K. Sandhoff, Chapter 14 sphingolipids: Metabolism and cell A. Lawlor, L.A. Lotta, M. Mangino, L. Le Marchand, E. Mathé, C.E. Matthews,
signaling, in: New Compr. Biochem, Elsevier, 2002, pp. 373–407, https://doi.org/ C. Menni, L.A. Mucci, R. Murphy, M. Oresic, E. Orwoll, J. Ose, A.C. Pereira, M.
10.1016/S0167-7306(02)36016-2. C. Playdon, L. Poston, J. Price, Q. Qi, K. Rexrode, A. Risch, J. Sampson, W.
[24] K. Furukawa, Y. Ohkawa, Y. Yamauchi, K. Hamamura, Y. Ohmi, K. Furukawa, J. Seow, H.D. Sesso, S.H. Shah, X.-O. Shu, G.C.S. Smith, U. Sovio, L. Stevens,
Fine tuning of cell signals by glycosylation, J. Biochem. (Tokyo) 151 (2012) R. Stolzenberg-Solomon, T. Takebayashi, T. Tillin, R. Travis, I. Tzoulaki, C.
573–578, https://doi.org/10.1093/jb/mvs043. M. Ulrich, R.S. Vasan, M. Verma, Y. Wang, N.J. Wareham, A. Wong, N. Younes,
[25] G. D’Angelo, S. Capasso, L. Sticco, D. Russo, Glycosphingolipids: synthesis and H. Zhao, W. Zheng, S.C. Moore, The consortium of metabolomics studies
functions, FEBS J. 280 (2013) 6338–6353, https://doi.org/10.1111/febs.12559. (COMETS): metabolomics in 47 prospective cohort studies, Am. J. Epidemiol. 188
[26] G. van Meer, D.R. Voelker, G.W. Feigenson, Membrane lipids: where they are and (2019) 991–1012, https://doi.org/10.1093/aje/kwz028.
how they behave, Nat. Rev. Mol. Cell Biol. 9 (2008) 112–124, https://doi.org/ [49] P. Narayanaswamy, S. Shinde, R. Sulc, R. Kraut, G. Staples, C.H. Thiam,
10.1038/nrm2330. R. Grimm, B. Sellergren, F. Torta, M.R. Wenk, Lipidomic “deep profiling”: an
[27] K. Simons, W.L.C. Vaz, Model systems, lipid rafts, and cell membranes, Annu. enhanced workflow to reveal new molecular species of signaling lipids, Anal.
Rev. Biophys. Biomol. Struct. 33 (2004) 269–295, https://doi.org/10.1146/ Chem. 86 (2014) 3043–3047, https://doi.org/10.1021/ac4039652.
annurev.biophys.32.110601.141803. [50] O. Quehenberger, A.M. Armando, A.H. Brown, S.B. Milne, D.S. Myers, A.
[28] K. Hanada, K. Kumagai, N. Tomishige, T. Yamaji, CERT-mediated trafficking of H. Merrill, S. Bandyopadhyay, K.N. Jones, S. Kelly, R.L. Shaner, C.M. Sullards,
ceramide, Biochim. Biophys. Acta BBA - Mol. Cell Biol. Lipids 1791 (2009) E. Wang, R.C. Murphy, R.M. Barkley, T.J. Leiker, C.R.H. Raetz, Z. Guan, G.
684–691, https://doi.org/10.1016/j.bbalip.2009.01.006. M. Laird, D.A. Six, D.W. Russell, J.G. McDonald, S. Subramaniam, E. Fahy, E.
[29] K. Huitema, J. van den Dikkenberg, J.F.H.M. Brouwers, J.C.M. Holthuis, A. Dennis, Lipidomics reveals a remarkable diversity of lipids in human plasma,
Identification of a family of animal sphingomyelin synthases, EMBO J. 23 (2004) J. Lipid Res. 51 (2010) 3299–3305, https://doi.org/10.1194/jlr.M009449.
33–44, https://doi.org/10.1038/sj.emboj.7600034. [51] B.J. Jenkins, K. Seyssel, S. Chiu, P.-H. Pan, S.-Y. Lin, E. Stanley, Z. Ament, J.
[30] P.J. Slotte, Molecular properties of various structurally defined sphingomyelins – A. West, K. Summerhill, J.L. Griffin, W. Vetter, K.J. Autio, K. Hiltunen,
correlation of structure with function, Prog. Lipid Res. 52 (2013) 206–219, S. Hazebrouck, R. Stepankova, C.-J. Chen, M. Alligier, M. Laville, M. Moore,
https://doi.org/10.1016/j.plipres.2012.12.001. G. Kraft, A. Cherrington, S. King, R.M. Krauss, E. de Schryver, P.P. Van
[31] O. Engberg, K.-L. Lin, V. Hautala, J.P. Slotte, T.K.M. Nyholm, Sphingomyelin acyl Veldhoven, M. Ronis, A. Koulman, Odd chain fatty acids; new insights of the
chains influence the formation of sphingomyelin- and cholesterol-enriched relationship between the gut microbiota, dietary intake, biosynthesis and glucose
domains, Biophys. J. 119 (2020) 913–923, https://doi.org/10.1016/j. intolerance, Sci. Rep. 7 (2017) 44845, https://doi.org/10.1038/srep44845.
bpj.2020.07.014. [52] B. Jenkins, J.A. West, A. Koulman, A review of odd-chain fatty acid metabolism
[32] A.E. Garner, D.A. Smith, N.M. Hooper, Sphingomyelin chain length influences the and the role of pentadecanoic acid (c15:0) and heptadecanoic acid (c17:0) in
distribution of GPI-anchored proteins in rafts in supported lipid bilayers, Mol. health and disease, Mol. Basel Switz. 20 (2015) 2425–2444, https://doi.org/
Membr. Biol. 24 (2007) 233–242, https://doi.org/10.1080/ 10.3390/molecules20022425.
09687860601127770.

9
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

[53] T. Hornemann, A. Penno, M.F. Rütti, D. Ernst, F. Kivrak-Pfiffner, L. Rohrer, A. von [74] W. Wang, P. Xiang, W.S. Chew, F. Torta, A. Bandla, V. Lopez, W.L. Seow, B.W.
Eckardstein, The SPTLC3 subunit of serine palmitoyltransferase generates short S. Lam, J.K. Chang, P. Wong, K. Chayaburakul, W.-Y. Ong, M.R. Wenk, R. Sundar,
chain Sphingoid bases, J. Biol. Chem. 284 (2009) 26322–26330, https://doi.org/ D.R. Herr, Activation of sphingosine 1-phosphate receptor 2 attenuates
10.1074/jbc.M109.023192. chemotherapy-induced neuropathy, J. Biol. Chem. 295 (2020) 1143–1152,
[54] M.A. Lone, A.J. Hülsmeier, E.M. Saied, G. Karsai, C. Arenz, A. von Eckardstein, https://doi.org/10.1074/jbc.RA119.011699.
T. Hornemann, Subunit composition of the mammalian serine- [75] T. Doyle, Z. Chen, L.M. Obeid, D. Salvemini, Sphingosine-1-phosphate acting via
palmitoyltransferase defines the spectrum of straight and methyl-branched long- the S1P₁ receptor is a downstream signaling pathway in ceramide-induced
chain bases, Proc. Natl. Acad. Sci. 117 (2020) 15591–15598, https://doi.org/ hyperalgesia, Neurosci. Lett. 499 (2011) 4–8, https://doi.org/10.1016/j.
10.1073/pnas.2002391117. neulet.2011.05.018.
[55] B.M. Castro, M. Prieto, L.C. Silva, Ceramide: a simple sphingolipid with unique [76] K. Janes, J.W. Little, C. Li, L. Bryant, C. Chen, Z. Chen, K. Kamocki, T. Doyle,
biophysical properties, Prog. Lipid Res. 54 (2014) 53–67, https://doi.org/ A. Snider, E. Esposito, S. Cuzzocrea, E. Bieberich, L. Obeid, I. Petrache, G. Nicol,
10.1016/j.plipres.2014.01.004. W.L. Neumann, D. Salvemini, The development and maintenance of paclitaxel-
[56] P.S. Niemelä, M.T. Hyvönen, I. Vattulainen, Influence of chain length and induced neuropathic pain require activation of the sphingosine 1-phosphate
unsaturation on sphingomyelin bilayers, Biophys. J. 90 (2006) 851–863, https:// receptor subtype 1, J. Biol. Chem. 289 (2014) 21082–21097, https://doi.org/
doi.org/10.1529/biophysj.105.067371. 10.1074/jbc.M114.569574.
[57] S.N. Pinto, E.L. Laviad, J. Stiban, S.L. Kelly, A.H. Merrill, M. Prieto, A. [77] K. Stockstill, T.M. Doyle, X. Yan, Z. Chen, K. Janes, J.W. Little, K. Braden,
H. Futerman, L.C. Silva, Changes in membrane biophysical properties induced by F. Lauro, L.A. Giancotti, C.M. Harada, R. Yadav, W.H. Xiao, J.M. Lionberger, W.
sphingomyelinase depend on the sphingolipid N-acyl chain, J. Lipid Res. 55 L. Neumann, G.J. Bennett, H.-R. Weng, S. Spiegel, D. Salvemini, Dysregulation of
(2014) 53–61, https://doi.org/10.1194/jlr.M042002. sphingolipid metabolism contributes to bortezomib-induced neuropathic pain,
[58] T. Maula, I. Artetxe, P.-M. Grandell, J.P. Slotte, Importance of the sphingoid base J. Exp. Med. 215 (2018) 1301–1313, https://doi.org/10.1084/jem.20170584.
length for the membrane properties of ceramides, Biophys. J. 103 (2012) [78] T.M. Doyle, K. Janes, Z. Chen, P.M. Grace, E. Esposito, S. Cuzzocrea, T.
1870–1879, https://doi.org/10.1016/j.bpj.2012.09.018. M. Largent-Milnes, W.L. Neumann, L.R. Watkins, S. Spiegel, T.W. Vanderah,
[59] M.A. Al Sazzad, T. Yasuda, M. Murata, J.P. Slotte, The long-chain sphingoid base D. Salvemini, Activation of sphingosine-1-phosphate receptor subtype 1 in the
of ceramides determines their propensity for lateral segregation, Biophys. J. 112 central nervous system contributes to morphine-induced hyperalgesia and
(2017) 976–983, https://doi.org/10.1016/j.bpj.2017.01.016. antinociceptive tolerance in rodents, Pain (2020), https://doi.org/10.1097/j.
[60] P.H.H. Lopez, R.L. Schnaar, Gangliosides in cell recognition and membrane pain.0000000000001888.
protein regulation, Curr. Opin. Struct. Biol. 19 (2009) 549–557, https://doi.org/ [79] D.A. Vessey, L. Li, M. Kelley, J. Zhang, J.S. Karliner, Sphingosine can pre- and
10.1016/j.sbi.2009.06.001. post-condition heart and utilizes a different mechanism from sphingosine 1-
[61] E. Di Biase, G. Lunghi, M. Fazzari, M. Maggioni, D.Y. Pomè, M. Valsecchi, phosphate, J. Biochem. Mol. Toxicol. 22 (2008) 113–118, https://doi.org/
M. Samarani, P. Fato, M.G. Ciampa, S. Prioni, L. Mauri, S. Sonnino, E. Chiricozzi, 10.1002/jbt.20227.
Gangliosides in the differentiation process of primary neurons: the specific role of [80] O. Cuvillier, V.E. Nava, S.K. Murthy, L.C. Edsall, T. Levade, S. Milstien, S. Spiegel,
GM1-oligosaccharide, Glycoconj. J. 37 (2020) 329–343, https://doi.org/ Sphingosine generation, cytochrome c release, and activation of caspase-7 in
10.1007/s10719-020-09919-x. doxorubicin-induced apoptosis of MCF7 breast adenocarcinoma cells, Cell Death
[62] E. Chiricozzi, E.D. Biase, M. Maggioni, G. Lunghi, M. Fazzari, D.Y. Pomè, Differ. 8 (2001) 162–171, https://doi.org/10.1038/sj.cdd.4400793.
R. Casellato, N. Loberto, L. Mauri, S. Sonnino, GM 1 promotes TrkA-mediated [81] K.A. Krown, M.T. Page, C. Nguyen, D. Zechner, V. Gutierrez, K.L. Comstock, C.
neuroblastoma cell differentiation by occupying a plasma membrane domain C. Glembotski, P.J. Quintana, R.A. Sabbadini, Tumor necrosis factor alpha-
different from TrkA, J. Neurochem. 149 (2019) 231–241, https://doi.org/ induced apoptosis in cardiac myocytes. Involvement of the sphingolipid signaling
10.1111/jnc.14685. cascade in cardiac cell death, J. Clin. Invest. 98 (1996) 2854–2865, https://doi.
[63] E. Chiricozzi, M. Maggioni, E. di Biase, G. Lunghi, M. Fazzari, N. Loberto, org/10.1172/JCI119114.
M. Elisa, F.G. Scalvini, G. Tedeschi, S. Sonnino, The neuroprotective role of the [82] E.A. Sweeney, J. Inokuchi, Y. Igarashi, Inhibition of sphingolipid induced
GM1 oligosaccharide, II3Neu5Ac-Gg4, in neuroblastoma cells, Mol. Neurobiol. 56 apoptosis by caspase inhibitors indicates that sphingosine acts in an earlier part of
(2019) 6673–6702, https://doi.org/10.1007/s12035-019-1556-8. the apoptotic pathway than ceramide, FEBS Lett. 425 (1998) 61–65, https://doi.
[64] E. Chiricozzi, D.Y. Pomè, M. Maggioni, E. Di Biase, C. Parravicini, L. Palazzolo, org/10.1016/S0014-5793(98)00198-7.
N. Loberto, I. Eberini, S. Sonnino, Role of the GM1 ganglioside oligosaccharide [83] A. Othman, C.H. Saely, A. Muendlein, A. Vonbank, H. Drexel, A. von Eckardstein,
portion in the TrkA-dependent neurite sprouting in neuroblastoma cells, T. Hornemann, Plasma C20-sphingolipids predict cardiovascular events
J. Neurochem. 143 (2017) 645–659, https://doi.org/10.1111/jnc.14146. independently from conventional cardiovascular risk factors in patients
[65] M. Fazzari, M. Audano, G. Lunghi, E. Di Biase, N. Loberto, L. Mauri, N. Mitro, undergoing coronary angiography, Atherosclerosis 240 (2015) 216–221, https://
S. Sonnino, E. Chiricozzi, The oligosaccharide portion of ganglioside GM1 doi.org/10.1016/j.atherosclerosis.2015.03.011.
regulates mitochondrial function in neuroblastoma cells, Glycoconj. J. 37 (2020) [84] Q. You, Q. Peng, Z. Yu, H. Jin, J. Zhang, W. Sun, Y. Huang, Plasma lipidomic
293–306, https://doi.org/10.1007/s10719-020-09920-4. analysis of sphingolipids in patients with large artery atherosclerosis
[66] S. Sonnino, V. Chigorno, Ganglioside molecular species containing C18- and C20- cerebrovascular disease and cerebral small vessel disease, Biosci. Rep. 40 (2020),
sphingosine in mammalian nervous tissues and neuronal cell cultures, Biochim. https://doi.org/10.1042/BSR20201519. BSR20201519.
Biophys. Acta Rev. Biomembr. 1469 (2000) 63–77, https://doi.org/10.1016/ [85] S.B. Russo, R. Tidhar, A.H. Futerman, L.A. Cowart, Myristate-derived d16:
S0005-2736(00)00210-8. 0 sphingolipids constitute a cardiac sphingolipid pool with distinct synthetic
[67] N. Weishaupt, S. Caughlin, K.K.-C. Yeung, S.N. Whitehead, Differential routes and functional properties, J. Biol. Chem. 288 (2013) 13397–13409,
anatomical expression of ganglioside GM1 species containing d18:1 or d20:1 https://doi.org/10.1074/jbc.M112.428185.
sphingosine detected by MALDI imaging mass spectrometry in mature rat brain, [86] W.L. Holland, B.T. Bikman, L.-P. Wang, G. Yuguang, K.M. Sargent, S. Bulchand, T.
Front. Neuroanat. 9 (2015), https://doi.org/10.3389/fnana.2015.00155. A. Knotts, G. Shui, D.J. Clegg, M.R. Wenk, M.J. Pagliassotti, P.E. Scherer, S.
[68] Y. Sugiura, S. Shimma, Y. Konishi, M.K. Yamada, M. Setou, Imaging mass A. Summers, Lipid-induced insulin resistance mediated by the proinflammatory
spectrometry technology and application on ganglioside study; visualization of receptor TLR4 requires saturated fatty acid-induced ceramide biosynthesis in
age-dependent accumulation of C20-ganglioside molecular species in the mouse mice, J. Clin. Invest. 121 (2011) 1858–1870, https://doi.org/10.1172/JCI43378.
hippocampus, PLoS One 3 (2008), e3232, https://doi.org/10.1371/journal. [87] K. Kurek, D.M. Piotrowska, P. Wiesiołek-Kurek, B. Łukaszuk, A. Chabowski,
pone.0003232. J. Górski, M. Zendzian-Piotrowska, Inhibition of ceramide de novo synthesis
[69] S. Caughlin, S. Maheshwari, N. Weishaupt, K.K.-C. Yeung, D.F. Cechetto, S. reduces liver lipid accumulation in rats with nonalcoholic fatty liver disease,
N. Whitehead, Age-dependent and regional heterogeneity in the long-chain base Liver Int. Off. J. Int. Assoc. Study Liver. 34 (2014) 1074–1083, https://doi.org/
of A-series gangliosides observed in the rat brain using MALDI imaging, Sci. Rep. 10.1111/liv.12331.
7 (2017) 16135, https://doi.org/10.1038/s41598-017-16389-z. [88] W. Teng, Y. Li, M. Du, X. Lei, S. Xie, F. Ren, Sulforaphane prevents hepatic insulin
[70] P. Palestini, S. Sonnino, G. Tettamanti, Lack of the ganglioside molecular species resistance by blocking serine palmitoyltransferase 3-mediated ceramide
containing the C20-long-chain bases in human, rat, mouse, rabbit, cat, dog, and biosynthesis, Nutrients 11 (2019), https://doi.org/10.3390/nu11051185.
chicken brains during prenatal life, J. Neurochem. 56 (1991) 2048–2050, https:// [89] H. Fyrst, B. Oskouian, P. Bandhuvula, Y. Gong, H.S. Byun, R. Bittman, A.R. Lee, J.
doi.org/10.1111/j.1471-4159.1991.tb03465.x. D. Saba, Natural sphingadienes inhibit Akt-dependent signaling and prevent
[71] X.Y. Chua, Y.L. Chai, W.S. Chew, J.R. Chong, H.L. Ang, P. Xiang, K. Camara, A. intestinal tumorigenesis, Cancer Res. 69 (2009) 9457–9464, https://doi.org/
R. Howell, F. Torta, M.R. Wenk, S. Hilal, N. Venketasubramanian, C.P. Chen, D. 10.1158/0008-5472.CAN-09-2341.
R. Herr, M.K.P. Lai, Immunomodulatory sphingosine-1-phosphates as plasma [90] H. Fyrst, X. Zhang, D.R. Herr, H.S. Byun, R. Bittman, V.H. Phan, G.L. Harris, J.
biomarkers of Alzheimer’s disease and vascular cognitive impairment, D. Saba, Identification and characterization by electrospray mass spectrometry of
Alzheimers Res. Ther. 12 (2020) 122, https://doi.org/10.1186/s13195-020- endogenous drosophila sphingadienes, J. Lipid Res. 49 (2008) 597–606, https://
00694-3. doi.org/10.1194/jlr.M700414-JLR200.
[72] R. Vutukuri, A. Koch, S. Trautmann, Y. Schreiber, D. Thomas, F. Mayser, D. Meyer [91] S.N. Pinto, L.C. Silva, A.H. Futerman, M. Prieto, Effect of ceramide structure on
Zu, J. Heringdorf, W. Pfeilschifter, R. Brunkhorst Pfeilschifter, S1P d20:1, an membrane biophysical properties: the role of acyl chain length and unsaturation,
endogenous modulator of S1P d18:1/S1P2 -dependent signaling, FASEB J. Off. Biochim. Biophys. Acta Biomembr. 1808 (2011) 2753–2760, https://doi.org/
Publ. Fed. Am. Soc. Exp. Biol. 34 (2020) 3932–3942, https://doi.org/10.1096/ 10.1016/j.bbamem.2011.07.023.
fj.201902391R. [92] T. Maula, M.A. Al Sazzad, J.P. Slotte, Influence of hydroxylation, chain length,
[73] A. Troupiotis-Tsaïlaki, J. Zachmann, I. González-Gil, A. Gonzalez, S. Ortega- and chain unsaturation on bilayer properties of ceramides, Biophys. J. 109 (2015)
Gutiérrez, M.L. López-Rodríguez, L. Pardo, C. Govaerts, Ligand chain length 1639–1651, https://doi.org/10.1016/j.bpj.2015.08.040.
drives activation of lipid G protein-coupled receptors, Sci. Rep. 7 (2017) 2020,
https://doi.org/10.1038/s41598-017-02104-5.

10
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

[93] M. Chakraborty, X.-C. Jiang, Sphingomyelin and its role in cellular signaling, in: [113] J.L. Dawkins, D.J. Hulme, S.B. Brahmbhatt, M. Auer-Grumbach, G.A. Nicholson,
D.G.S. Capelluto (Ed.), Lipid-Mediat. Protein Signal, Springer Netherlands, Mutations in SPTLC1, encoding serine palmitoyltransferase, long chain base
Dordrecht, 2013, pp. 1–14, https://doi.org/10.1007/978-94-007-6331-9_1. subunit-1, cause hereditary sensory neuropathy type I, Nat. Genet. 27 (2001)
[94] R.V. Panganamala, J.C. Geer, D.G. Cornwell, Long-chain bases in the 309–312, https://doi.org/10.1038/85879.
sphingolipids of atherosclerotic human aorta, J. Lipid Res. 10 (1969) 445–455. [114] J. Ren, E.M. Saied, A. Zhong, J. Snider, C. Ruiz, C. Arenz, L.M. Obeid, G.
[95] O. Renkonen, E.L. Hirvisalo, Structure of plasma sphingadienine, J. Lipid Res. 10 D. Girnun, Y.A. Hannun, Tsc3 regulates SPT amino acid choice in Saccharomyces
(1969) 687–693. cerevisiae by promoting alanine in the sphingolipid pathway, J. Lipid Res. 59
[96] G. Karsai, M. Lone, Z. Kutalik, J.T. Brenna, H. Li, D. Pan, A. von Eckardstein, (2018) 2126–2139, https://doi.org/10.1194/jlr.M088195.
T. Hornemann, FADS3 is a Δ14Z sphingoid base desaturase that contributes to [115] K. Bejaoui, Y. Uchida, S. Yasuda, M. Ho, M. Nishijima, R.H. Brown, W.
gender differences in the human plasma sphingolipidome, J. Biol. Chem. 295 M. Holleran, K. Hanada, Hereditary sensory neuropathy type 1 mutations confer
(2020) 1889–1897, https://doi.org/10.1074/jbc.AC119.011883. dominant negative effects on serine palmitoyltransferase, critical for sphingolipid
[97] K. Jojima, M. Edagawa, M. Sawai, Y. Ohno, A. Kihara, Biosynthesis of the anti- synthesis, J. Clin. Invest. 110 (2002) 1301–1308, https://doi.org/10.1172/
lipid-microdomain sphingoid base 4,14-sphingadiene by the ceramide desaturase JCI0216450.
FADS3, FASEB J. 34 (2020) 3318–3335, https://doi.org/10.1096/ [116] A. Rotthier, M. Auer-Grumbach, K. Janssens, J. Baets, A. Penno, L. Almeida-
fj.201902645R. Souza, K. Van Hoof, A. Jacobs, E. De Vriendt, B. Schlotter-Weigel, W. Löscher,
[98] D.R. Adams, S. Pyne, N.J. Pyne, Structure-function analysis of lipid substrates and P. Vondráček, P. Seeman, P. De Jonghe, P. Van Dijck, A. Jordanova,
inhibitors of sphingosine kinases, Cell. Signal. 109806 (2020), https://doi.org/ T. Hornemann, V. Timmerman, Mutations in the SPTLC2 subunit of serine
10.1016/j.cellsig.2020.109806. palmitoyltransferase cause hereditary sensory and autonomic neuropathy type I,
[99] T.A. Couttas, Y.H. Rustam, H. Song, Y. Qi, J.D. Teo, J. Chen, G.E. Reid, A.S. Don, Am. J. Hum. Genet. 87 (2010) 513–522, https://doi.org/10.1016/j.
A novel function of sphingosine kinase 2 in the metabolism of sphinga-4,14-diene ajhg.2010.09.010.
lipids, Metabolites 10 (2020), https://doi.org/10.3390/metabo10060236. [117] R. Steiner, E.M. Saied, A. Othman, C. Arenz, A.T. Maccarone, B.L.J. Poad, S.
[100] H. Symolon, E.M. Schmelz, D.L. Dillehay, A.H. Merrill, Dietary soy sphingolipids J. Blanksby, A. von Eckardstein, T. Hornemann, Elucidating the chemical
suppress tumorigenesis and gene expression in 1,2-dimethylhydrazine-treated structure of native 1-deoxysphingosine, J. Lipid Res. 57 (2016) 1194–1203,
CF1 mice and ApcMin/+ mice, J. Nutr. 134 (2004) 1157–1161, https://doi.org/ https://doi.org/10.1194/jlr.M067033.
10.1093/jn/134.5.1157. [118] R. Mashima, T. Okuyama, M. Ohira, Biosynthesis of long chain base in
[101] A. Kumar, A.K. Pandurangan, F. Lu, H. Fyrst, M. Zhang, H.-S. Byun, R. Bittman, J. sphingolipids in animals, plants and fungi, Future Sci. OA 6 (2020), https://doi.
D. Saba, Chemopreventive sphingadienes downregulate Wnt signaling via a org/10.2144/fsoa-2019-0094 (n.d.).
PP2A/Akt/GSK3β pathway in colon cancer, Carcinogenesis 33 (2012) [119] Y. Mizutani, A. Kihara, Y. Igarashi, Identification of the human sphingolipid C4-
1726–1735, https://doi.org/10.1093/carcin/bgs174. hydroxylase, hDES2, and its up-regulation during keratinocyte differentiation,
[102] P. Zhao, A.E. Aguilar, J.Y. Lee, L.A. Paul, J.H. Suh, L. Puri, M. Zhang, FEBS Lett. 563 (2004) 93–97, https://doi.org/10.1016/S0014-5793(04)00274-1.
J. Beckstead, A. Witkowski, R.O. Ryan, J.D. Saba, Sphingadienes show [120] E.R. Wilson, U. Kugathasan, A.Y. Abramov, A.J. Clark, D.L.H. Bennett, M.
therapeutic efficacy in neuroblastoma in vitro and in vivo by targeting the AKT M. Reilly, L. Greensmith, B. Kalmar, Hereditary sensory neuropathy type 1-asso­
signaling pathway, Investig. New Drugs 36 (2018) 743–754, https://doi.org/ ciated deoxysphingolipids cause neurotoxicity, acute calcium handling
10.1007/s10637-017-0558-5. abnormalities and mitochondrial dysfunction in vitro, Neurobiol. Dis. 117 (2018)
[103] J.H. Suh, A.M. Makarova, J.M. Gomez, L.A. Paul, J.D. Saba, An LC/MS/MS 1–14, https://doi.org/10.1016/j.nbd.2018.05.008.
method for quantitation of chemopreventive sphingadienes in food products and [121] A.G. Haribowo, J.T. Hannich, A.H. Michel, M. Megyeri, M. Schuldiner,
biological samples, J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 1061–1062 B. Kornmann, H. Riezman, Cytotoxicity of 1-deoxysphingolipid unraveled by
(2017) 292–299, https://doi.org/10.1016/j.jchromb.2017.07.040. genome-wide genetic screens and lipidomics in Saccharomyces cerevisiae, Mol.
[104] A. Fujii, Y. Manabe, K. Aida, T. Tsuduki, T. Hirata, T. Sugawara, Selective Biol. Cell 30 (2019) 2814–2826, https://doi.org/10.1091/mbc.E19-07-0364.
absorption of dietary sphingoid bases from the intestine via efflux by P- [122] I. Alecu, A. Tedeschi, N. Behler, K. Wunderling, C. Lamberz, M.A.R. Lauterbach,
glycoprotein in rats, J. Nutr. Sci. Vitaminol. (Tokyo) 63 (2017) 44–50, https:// A. Gaebler, D. Ernst, P.P. Van Veldhoven, A. Al-Amoudi, E. Latz, A. Othman,
doi.org/10.3177/jnsv.63.44. L. Kuerschner, T. Hornemann, F. Bradke, C. Thiele, A. Penno, Localization of 1-
[105] S. Usuki, N. Tamura, T. Tamura, K. Tanji, D. Mikami, K. Mukai, Y. Igarashi, deoxysphingolipids to mitochondria induces mitochondrial dysfunction, J. Lipid
Neurite outgrowth and morphological changes induced by 8-trans unsaturation of Res. 58 (2017) 42–59, https://doi.org/10.1194/jlr.M068676.
sphingadienine in kCer molecular species, Int. J. Mol. Sci. 20 (2019), https://doi. [123] R. Kramer, J. Bielawski, E. Kistner-Griffin, A. Othman, I. Alecu, D. Ernst,
org/10.3390/ijms20092116. E. Kornhauser, T. Hornemann, S. Spassieva, Neurotoxic 1-deoxysphingolipids and
[106] N.C. Zitomer, T. Mitchell, K.A. Voss, G.S. Bondy, S.T. Pruett, E.C. Garnier- paclitaxel-induced peripheral neuropathy, FASEB J. 29 (2015) 4461–4472,
Amblard, L.S. Liebeskind, H. Park, E. Wang, M.C. Sullards, A.H. Merrill, R. https://doi.org/10.1096/fj.15-272567.
T. Riley, Ceramide synthase inhibition by fumonisin B 1 causes accumulation of 1- [124] K.A. Becker, A.-K. Uerschels, L. Goins, S. Doolen, K.J. McQuerry, J. Bielawski,
deoxysphinganine: a novel category of bioactive 1-deoxysphingoid bases and 1- U. Sure, E. Bieberich, B.K. Taylor, E. Gulbins, S.D. Spassieva, Role of 1-Deoxy­
deoxydihydroceramides biosynthesized by mammalian cell lines and animals, sphingolipids in docetaxel neurotoxicity, J. Neurochem. 154 (2020) 662–672,
J. Biol. Chem. 284 (2009) 4786–4795, https://doi.org/10.1074/jbc. https://doi.org/10.1111/jnc.14985.
M808798200. [125] H.-S. Byun, S. Pyne, N. Macritchie, N.J. Pyne, R. Bittman, Novel sphingosine-
[107] A. Othman, M.F. Rütti, D. Ernst, C.H. Saely, P. Rein, H. Drexel, C. Porretta- containing analogues selectively inhibit sphingosine kinase (SK) isozymes, induce
Serapiglia, G. Lauria, R. Bianchi, A. von Eckardstein, T. Hornemann, Plasma SK1 proteasomal degradation and reduce DNA synthesis in human pulmonary
deoxysphingolipids: a novel class of biomarkers for the metabolic syndrome? arterial smooth muscle cells, MedChemComm 4 (2013), https://doi.org/10.1039/
Diabetologia 55 (2012) 421–431, https://doi.org/10.1007/s00125-011-2384-1. C3MD00201B.
[108] H.U. Humpf, E.M. Schmelz, F.I. Meredith, H. Vesper, T.R. Vales, E. Wang, D. [126] N.J. Pyne, M. McNaughton, S. Boomkamp, N. MacRitchie, C. Evangelisti, A.
S. Menaldino, D.C. Liotta, A.H. Merrill, Acylation of naturally occurring and M. Martelli, H.-R. Jiang, S. Ubhi, S. Pyne, Role of sphingosine 1-phosphate
synthetic 1-deoxysphinganines by ceramide synthase. Formation of N-palmitoyl- receptors, sphingosine kinases and sphingosine in cancer and inflammation, Adv.
aminopentol produces a toxic metabolite of hydrolyzed fumonisin, AP1, and a Biol. Regul. 60 (2016) 151–159, https://doi.org/10.1016/j.jbior.2015.09.001.
new category of ceramide synthase inhibitor, J. Biol. Chem. 273 (1998) [127] J. Mwinyi, A. Boström, I. Fehrer, A. Othman, G. Waeber, H. Marti-Soler,
19060–19064, https://doi.org/10.1074/jbc.273.30.19060. P. Vollenweider, P. Marques-Vidal, H.B. Schiöth, A. von Eckardstein,
[109] K. Garofalo, A. Penno, B.P. Schmidt, H.-J. Lee, M.P. Frosch, A. von Eckardstein, R. T. Hornemann, Plasma 1-deoxysphingolipids are early predictors of incident type
H. Brown, T. Hornemann, F.S. Eichler, Oral l-serine supplementation reduces 2 diabetes mellitus, PLoS One 12 (2017), https://doi.org/10.1371/journal.
production of neurotoxic deoxysphingolipids in mice and humans with hereditary pone.0175776 e0175776.
sensory autonomic neuropathy type 1, J. Clin. Invest. 121 (2011) 4735–4745, [128] N. Wei, J. Pan, R. Pop-Busui, A. Othman, I. Alecu, T. Hornemann, F.S. Eichler,
https://doi.org/10.1172/JCI57549. Altered sphingoid base profiles in type 1 compared to type 2 diabetes, Lipids
[110] K. Esaki, T. Sayano, C. Sonoda, T. Akagi, T. Suzuki, T. Ogawa, M. Okamoto, Health Dis. 13 (2014) 161, https://doi.org/10.1186/1476-511X-13-161.
T. Yoshikawa, Y. Hirabayashi, S. Furuya, L-serine deficiency elicits intracellular [129] R.A. Zuellig, T. Hornemann, A. Othman, A.B. Hehl, H. Bode, T. Güntert, O.
accumulation of cytotoxic deoxysphingolipids and lipid body formation, J. Biol. O. Ogunshola, E. Saponara, K. Grabliauskaite, J.-H. Jang, U. Ungethuem, Y. Wei,
Chem. 290 (2015) 14595–14609, https://doi.org/10.1074/jbc.M114.603860. A. von Eckardstein, R. Graf, S. Sonda, Deoxysphingolipids, novel biomarkers for
[111] A. Penno, M.M. Reilly, H. Houlden, M. Laurá, K. Rentsch, V. Niederkofler, E. type 2 diabetes, are cytotoxic for insulin-producing cells, Diabetes 63 (2014)
T. Stoeckli, G. Nicholson, F. Eichler, R.H. Brown, A. von Eckardstein, 1326–1339, https://doi.org/10.2337/db13-1042.
T. Hornemann, Hereditary sensory neuropathy type 1 is caused by the [130] A. Kowluru, Deoxysphingolipids: β-cell, beware of these new kids on the block,
accumulation of two neurotoxic sphingolipids, J. Biol. Chem. 285 (2010) Diabetes 63 (2014) 1191–1193, https://doi.org/10.2337/db14-0022.
11178–11187, https://doi.org/10.1074/jbc.M109.092973. [131] M.F. Dohrn, A. Othman, S.K. Hirshman, H. Bode, I. Alecu, E. Fähndrich,
[112] M.L. Gantner, K. Eade, M. Wallace, M.K. Handzlik, R. Fallon, J. Trombley, W. Karges, J. Weis, J.B. Schulz, T. Hornemann, K.G. Claeys, Elevation of plasma
R. Bonelli, S. Giles, S. Harkins-Perry, T.F.C. Heeren, L. Sauer, Y. Ideguchi, 1-deoxy-sphingolipids in type 2 diabetes mellitus: a susceptibility to neuropathy?
M. Baldini, L. Scheppke, M.I. Dorrell, M. Kitano, B.J. Hart, C. Cai, T. Nagasaki, M. Eur. J. Neurol. 22 (2015) 806–e55, https://doi.org/10.1111/ene.12663.
G. Badur, M. Okada, S.M. Woods, C. Egan, M. Gillies, R. Guymer, F. Eichler, [132] T. Muthusamy, T. Cordes, M.K. Handzlik, L. You, E.W. Lim, J. Gengatharan, A.F.
M. Bahlo, M. Fruttiger, R. Allikmets, P.S. Bernstein, C.M. Metallo, M. Friedlander, M. Pinto, M.G. Badur, M.J. Kolar, M. Wallace, A. Saghatelian, C.M. Metallo,
Serine and lipid metabolism in macular disease and peripheral neuropathy, Serine restriction alters sphingolipid diversity to constrain tumour growth, Nature
N. Engl. J. Med. 381 (2019) 1422–1433, https://doi.org/10.1056/ 586 (2020) 790–795, https://doi.org/10.1038/s41586-020-2609-x.
NEJMoa1815111. [133] R.S. Banh, D.E. Biancur, K. Yamamoto, A.S.W. Sohn, B. Walters, M. Kuljanin,
A. Gikandi, H. Wang, J.D. Mancias, R.J. Schneider, M.E. Pacold, A.C. Kimmelman,

11
B.W.S. Lam et al. Cellular Signalling 79 (2021) 109890

Neurons release serine to support mRNA translation in pancreatic cancer, Cell 183 restore extracellular matrix in human dermal fibroblasts in vitro and in vivo, Int.
(2020) 1202–1218, e25, https://doi.org/10.1016/j.cell.2020.10.016. J. Mol. Med. 39 (2017) 741–748, https://doi.org/10.3892/ijmm.2017.2866.
[134] L. Coderch, O. López, A. de la Maza, J.L. Parra, Ceramides and skin function, Am. [145] Y.B. Lee, E.J. Byun, H.S. Kim, Potential role of the microbiome in acne: a
J. Clin. Dermatol. 4 (2003) 107–129, https://doi.org/10.2165/00128071- comprehensive review, J. Clin. Med. 8 (2019) 987, https://doi.org/10.3390/
200304020-00004. jcm8070987.
[135] H.K. Choi, H.-J. Kim, K.-H. Liu, C.S. Park, Phytosphingosine increases [146] K.-M. Park, J.W. Wang, Y.-M. Yoo, M.J. Choi, K.C. Hwang, E.-B. Jeung, Y.
biosynthesis of phytoceramide by uniquely stimulating the expression of W. Jeong, W.S. Hwang, Sphingosine-1-phosphate (S1P) analog phytosphingosine-
dihydroceramide C4-desaturase (DES2) in cultured human keratinocytes, Lipids 1-phosphate (P1P) improves the in vitro maturation efficiency of porcine oocytes
53 (2018) 909–918, https://doi.org/10.1002/lipd.12097. via regulation of oxidative stress and apoptosis, Mol. Reprod. Dev. 86 (2019)
[136] Ceramide composition of the psoriatic scale, Biochim. Biophys. Acta (BBA) - Mol. 1705–1719, https://doi.org/10.1002/mrd.23264.
Basis Dis. 1182 (1993) 147–151, https://doi.org/10.1016/0925-4439(93)90135- [147] H.-M. Kang, H.-S. Son, Y.-H. Cui, B. Youn, B. Son, N.K. Kaushik, N. Uddin, J.-
N. S. Lee, J.-Y. Song, N. Kaushik, S.-J. Lee, Phytosphingosine exhibits an anti-
[137] J. Ishikawa, H. Narita, N. Kondo, M. Hotta, Y. Takagi, Y. Masukawa, T. Kitahara, epithelial–mesenchymal transition function by the inhibition of EGFR signaling in
Y. Takema, S. Koyano, S. Yamazaki, A. Hatamochi, Changes in the ceramide human breast cancer cells, Oncotarget 8 (2017) 77794–77808, https://doi.org/
profile of atopic dermatitis patients, J. Invest. Dermatol. 130 (2010) 2511–2514, 10.18632/oncotarget.20783.
https://doi.org/10.1038/jid.2010.161. [148] E.J. Jang, Y. Shin, H.J. Park, D. Kim, C. Jung, J.-Y. Hong, S. Kim, S.K. Lee, Anti-
[138] K. Ishida, A. Takahashi, K. Bito, Z. Draelos, G. Imokawa, Treatment with synthetic melanogenic activity of phytosphingosine via the modulation of the
pseudoceramide improves atopic skin, switching the ceramide profile to a healthy microphthalmia-associated transcription factor signaling pathway, J. Dermatol.
skin phenotype, J. Invest. Dermatol. 140 (2020) 1762–1770, e8, https://doi. Sci. 87 (2017) 19–28, https://doi.org/10.1016/j.jdermsci.2017.03.011.
org/10.1016/j.jid.2020.01.014. [149] H.E. Seberg, E.V. Otterloo, R.A. Cornell, Beyond MITF: multiple transcription
[139] H.K. Choi, Y.H. Cho, E.O. Lee, J.W. Kim, C.S. Park, Phytosphingosine enhances factors directly regulate the cellular phenotype in melanocytes and melanoma,
moisture level in human skin barrier through stimulation of the filaggrin Pigment Cell Melanoma Res. 30 (2017) 454–466, https://doi.org/10.1111/
biosynthesis and degradation leading to NMF formation, Arch. Dermatol. Res. 309 pcmr.12611.
(2017) 795–803, https://doi.org/10.1007/s00403-017-1782-8. [150] Y. Gao, T. Xu, Y.-X. Zhao, T. Ling-Hu, S.-B. Liu, J.-S. Tian, X.-M. Qin, A. Novel
[140] Y. Inagaki, T.T. Pham, Y. Fujiwara, T. Kohno, D.A. Osborne, Y. Igarashi, G. Tigyi, Network, Pharmacology strategy to decode metabolic biomarkers and targets
A.L. Parrill, Sphingosine 1-phosphate analogue recognition and selectivity at interactions for depression, Front. Psychiatry 11 (2020), https://doi.org/
S1P4 within the endothelial differentiation gene family of receptors, Biochem. J. 10.3389/fpsyt.2020.00667.
389 (2005) 187–195, https://doi.org/10.1042/BJ20050046. [151] J.F. Cryan, K.J. O’Riordan, C.S.M. Cowan, K.V. Sandhu, T.F.S. Bastiaanssen,
[141] M. Rios Candelore, M.J. Wright, L.M. Tota, J. Milligan, G. Shei, J.D. Bergstrom, S. M. Boehme, M.G. Codagnone, S. Cussotto, C. Fulling, A.V. Golubeva, K.
M. Mandala, Phytosphingosine 1-phosphate: a high affinity ligand for the S1P4/ E. Guzzetta, M. Jaggar, C.M. Long-Smith, J.M. Lyte, J.A. Martin, A. Molinero-
Edg-6 receptor, Biochem. Biophys. Res. Commun. 297 (2002) 600–606, https:// Perez, G. Moloney, E. Morelli, E. Morillas, R. O’Connor, J.S. Cruz-Pereira, V.
doi.org/10.1016/S0006-291X(02)02237-4. L. Peterson, K. Rea, N.L. Ritz, E. Sherwin, S. Spichak, E.M. Teichman, M. van de
[142] M.-K. Kim, K.S. Park, H. Lee, Y.D. Kim, J. Yun, Y.-S. Bae, Phytosphingosine-1- Wouw, A.P. Ventura-Silva, S.E. Wallace-Fitzsimons, N. Hyland, G. Clarke, T.
phosphate stimulates chemotactic migration of L2071 mouse fibroblasts via G. Dinan, The microbiota-gut-brain axis, Physiol. Rev. 99 (2019) 1877–2013,
pertussis toxin-sensitive G-proteins, Exp. Mol. Med. 39 (2007) 185–194, https:// https://doi.org/10.1152/physrev.00018.2018.
doi.org/10.1038/emm.2007.21. [152] K. Huynh, C.K. Barlow, K.S. Jayawardana, J.M. Weir, N.A. Mellett, M. Cinel, D.
[143] J.P. Lee, H.J. Cha, K.S. Lee, K.K. Lee, J.H. Son, K.N. Kim, D.K. Lee, S. An, J. Magliano, J.E. Shaw, B.G. Drew, P.J. Meikle, High-throughput plasma
Phytosphingosine-1-phosphate represses the hydrogen peroxide-induced lipidomics: detailed mapping of the associations with cardiometabolic risk
activation of c-Jun N-terminal kinase in human dermal fibroblasts through the factors, Cell Chem. Biol. 26 (2019) 71–84, e4, https://doi.org/10.1016/j.ch
phosphatidylinositol 3-kinase/Akt pathway, Arch. Dermatol. Res. 304 (2012) embiol.2018.10.008.
673–678, https://doi.org/10.1007/s00403-012-1241-5. [153] S. Motta, M. Monti, S. Sesana, R. Caputo, S. Carelli, R. Ghidoni, Ceramide
[144] S.B. Kwon, S. An, M.J. Kim, K.R. Kim, Y.M. Choi, K.J. Ahn, I.-S. An, H.J. Cha, composition of the psoriatic scale, Biochim. Biophys. Acta 1182 (1993) 147–151,
Phytosphingosine-1-phosphate and epidermal growth factor synergistically https://doi.org/10.1016/0925-4439(93)90135-n.

12

You might also like