Next Article in Journal
Utilising a Clinical Metabolomics LC-MS Study to Determine the Integrity of Biological Samples for Statistical Modelling after Long Term −80 °C Storage: A TOFI_Asia Sub-Study
Next Article in Special Issue
Is Lipid Metabolism of Value in Cancer Research and Treatment? Part II: Role of Specialized Pro-Resolving Mediators in Inflammation, Infections, and Cancer
Previous Article in Journal
Machine Learning Metabolomics Profiling of Dietary Interventions from a Six-Week Randomised Trial
Previous Article in Special Issue
The Metabolic and Lipidomic Fingerprint of Torin1 Exposure in Mouse Embryonic Fibroblasts Using Untargeted Metabolomics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Is Lipid Metabolism of Value in Cancer Research and Treatment? Part I- Lipid Metabolism in Cancer

1
Department of Immunobiology, Yale University, West Haven, CT 06516, USA
2
Sciex, 500 Old Connecticut Path, Framingham, MA 01701, USA
3
Department of Microbiology, Hyogo Medical University, Nishinomiya 663-8501, Japan
4
Department of Pathology, Yale University, New Haven, CT 06520, USA
5
Vascular Biology and Therapeutic Program, Yale University School of Medicine, New Haven, CT 06520, USA
*
Author to whom correspondence should be addressed.
Metabolites 2024, 14(6), 312; https://doi.org/10.3390/metabo14060312
Submission received: 15 March 2024 / Revised: 12 May 2024 / Accepted: 16 May 2024 / Published: 29 May 2024

Abstract

:
For either healthy or diseased organisms, lipids are key components for cellular membranes; they play important roles in numerous cellular processes including cell growth, proliferation, differentiation, energy storage and signaling. Exercise and disease development are examples of cellular environment alterations which produce changes in these networks. There are indications that alterations in lipid metabolism contribute to the development and progression of a variety of cancers. Measuring such alterations and understanding the pathways involved is critical to fully understand cellular metabolism. The demands for this information have led to the emergence of lipidomics, which enables the large-scale study of lipids using mass spectrometry (MS) techniques. Mass spectrometry has been widely used in lipidomics and allows us to analyze detailed lipid profiles of cancers. In this article, we discuss emerging strategies for lipidomics by mass spectrometry; targeted, as opposed to global, lipid analysis provides an exciting new alternative method. Additionally, we provide an introduction to lipidomics, lipid categories and their major biological functions, along with lipidomics studies by mass spectrometry in cancer samples. Further, we summarize the importance of lipid metabolism in oncology and tumor microenvironment, some of the challenges for lipodomics, and the potential for targeted approaches for screening pharmaceutical candidates to improve the therapeutic efficacy of treatment in cancer patients.

1. Introduction

Lipids, which are fatty mixtures of fats, waxes, sterols, fat-soluble vitamins (vitamins A, D, E and K, for example), monoglycerides, diglycerides, triglycerides and phospholipids, achieve a range of functions in our body. Those lipids, which are part of the cell membranes, serve to help control what goes in and out of our cells, as well as helping with moving and storing energy, absorbing vitamins and making hormones. Having an excessive amount of some lipids is harmful; a lipid panel assay can tell us if a person has normal levels. Hundreds of thousands of individual lipid molecular species are present in cells. Phospholipids represent the bulk of lipids in mammalian cells, accounting for about 60 mol% of the total, while glycolipids and sphingolipids combine to account for about 10 mol%. The distribution of non-polar lipids, among which are triacylglyceride (TAG) and cholesterol, varies among cell types and compartments, ranging from 0.1 to 40 mol%. Natural metabolites form within the cell, generally comprising less than 5 mol% of total lipids. This group includes non-esterified fatty acids (NEFAs, or free fatty acids), lysolipids, diacylglycerol, ceramides, acyl carnitines, and acyl CoA. Excessive levels can produce harmful pathophysiologic sequelae under a variety of pathologic conditions [1,2,3,4].
The lipidome represents the total lipid content in a cell; the variety of these lipids is predicted to be in the range of tens of thousands to hundreds of thousands, with a wide range of concentrations. There are several classes and subclasses of lipids, characterized by the head group and the type of linkage between the head group and aliphatic chains (Figure 1) [5].

2. Lipid Categories and Their Major Biological Functions [5,6,7,8,9,10,11]

Based on their chemical structure, lipids have been classified into eight categories, as shown in Table 1 [11]:
  • Fatty acyls: The molecules in fatty acyl groups are synthesized by chain elongation of acetyl coenzyme A (CoA) primers with malonyl–CoA groups. They may contain cyclic functional groups and/or be substituted by heteroatoms. The structure of fatty acyls is the simplest, as they are characterized by a repeating series of methylene groups. Given their simplicity, fatty acyls function as the basic components of other, more complex lipids.
  • Prenol lipids: Five-carbon isoprene units represent the basic building blocks for prenol lipids. They may contain linear, cyclic or branched isoprene chains, any of which may consist of fatty acids, characteristic side-chains, functional groups, sugars, and double bonds. Given the range of permutations, these basic structures may form thousands of unique lipids. Prenol lipids are based on one or multiple five-carbon isoprene units, the building blocks of prenol lipids. The isoprene chain can be linear, cyclic or branched and features fatty acids, characteristic side-chains, functional groups, sugars and double bonds. The flexibility of the prenol lipids’ structure gives rise to thousands of different lipid species.
  • Glycerolipids: The glycerolipid groups include monoacylglycerol, diacylglycerol, triacylglycerol (TAG) and glycolipids, and they are characterized by the presence of a glycerol backbone with fatty acyl chains connected to the hydroxyl groups of glycerol. Glycerolipids can be hydrolyzed into glycerol, non-esterified fatty acid (NEFA) and/or alkyl variants; they contribute to energy storage, energy metabolism and signal transduction.
  • Glycerophospholipids: Glycerophospholipids (GPLs) contain at least one glycerol hydroxyl group esterified with one phosphate or phosphonate group. This category includes phosphatidylcholine (PC), phosphatidyl ethanolamine (PE), phosphatidylinositol (PI), phosphatidylserine (PS), phosphatidylglycerol (PG), phosphatidic acid (PA), cardiolipin, bis(monoacylglycerol) phosphate and lysophospholipids, etc. GPLs represent the bulk of cellular and plasma membranes and are one of their key components. GPLs contribute to second messenger generation and are involved in cellular metabolism and signal transduction.
  • Sphingolipids: Sphingolipid groups act to control cellular signaling; they are components of plasma membranes. They are characterized by their core structure, specifically long-chain sphingoids such as ceramide (Cer), sphingomyelin, cerebroside, sulfatide and gangliosides.
  • Saccharolipids: The saccharolipids are a group of lipids which contain a sugar backbone directly connected with fatty acids (as opposed to the glycerol backbone in glycerolipids and glycerophospholipids). They exist of glycans or phosphorylated derivative forms.
  • Sterol lipids: Sterol lipids are compounds containing four fused rings. Among the subcategories of sterol lipids are cholesterol and its derivatives, steroids, bile acids and their derivatives, etc. Cholesterol, along with its derivatives, is a vital component of cellular membranes. Steroids function as hormones and signaling molecules and are important to a variety of additional biological and metabolic processes.
  • Polyketides: The polyketides are a group of metabolites with characteristics similar to lipids. They come from plant and microbial sources. This category of lipids functions as cellular membrane components, energy storage depots and signaling molecules (Table 1).
It is well known that lipid metabolites alter tumors; indeed, this is considered a hallmark of cancer. There is a growing body of evidence that tumor growth is promoted through lipid metabolism deregulation; furthermore, it aids in cell migration, invasion and angiogenesis. Table 2 presents a list of the major functions of individual lipid classes including membrane components, energy storage and metabolism, signaling and other functions such as having antioxidant activity, acting in lipid transport and mitochondrial respiration, and as a cofactor and substrate of PE synthesis. For example, cholesterol defines and regulates the function and extent of lipid rafts on membranes. Some glycerolipids, such as NEFA, acylcarnitine, acyl CoA, diacylglycerol and TAG, are involved in energy metabolism and act as energy storage depots. Table 2 provides a summary of lipids contributing to signaling regulation.

3. Lipidomics Studies in Cancer Samples (MS Tech)

3.1. Sample Preparation and Lipid Extraction

The wide and varied range of lipid classes precludes the use of a standard, one-size-fits-all extraction technique. A generalized experimental workflow for lipid analysis is shown in Figure 2 [12]. As always, prior to analysis, the samples must be prepared, stored and processed properly. Fast, reproducible preparation methods which are able to handle various lipids and their different polarities while maintaining compatibility with the instrumental technique are ideal. Three major techniques are liquid–liquid extraction, organic solvent precipitation and solid-phase extraction, with organic solvent precipitation methods being the most common; each offers high rates of lipid recovery for lipodomics researchers. An early standard, the Bligh and Dyer procedure, involves chloroform–methanol based protocols which produce phase partitioning into the organic layer. More recent protocols have improved this method, producing better targeted lipid isolation and increased high-throughput data collection. While these methods work well for physiologically relevant lipids, one remaining problem area involves analysis of low-abundance and labile lipid metabolites [12,13,14,15,16,17].
The common extraction methods used in lipidomics
  • Modified Bligh and Dyer. Chloroform/methanol/H2O (1:1:0.9, v:v:v) for extraction of a small amount of biological sample (e.g., <50 mg of tissue). It is well-established and broadly practiced. The disadvantages include the use of hazardous chloroform and the collection of chloroform extract from the bottom layer, which may cause the carry-over of water-soluble impurities and difficulty in automation.
  • Modified Folch. Chloroform/methanol (2:1, v:v), 0.9% NaCl (0.2 volume) is used to extract biological tissue (e.g., ~0.1 g). Water or 0.9% NaCl (0.2 volume) is then added to wash the solvent extract. It has similar advantages and disadvantages to the modified Bligh and Dyer method.
  • Methyl tert-butyl ether (MTBE) method. The ratios are MTBE/methanol/water (5:1.5:1.45, v:v:v). This method resolves some of the difficulties in chloroform-involved methods because MTBE is present in the top layer after phase separation and therefore is more feasible for high throughput and automation. A disadvantage is that the MTBE phase contains a significant amount of aqueous components that may produce carry-over of water-soluble contaminants.
  • Butanol/methanol method. Start with a volume of butanol/methanol (BUME, 3:1, v/v) added to a small volume of aqueous phase. Next, an equal volume of heptane/ethyl acetate (3:1, v/v) is added, and then 1% acetic acid is added (equal volume to BUME) to induce phase separation. This method may compensate for the issue with the previous method, with less water-soluble contaminants being carried over in the organic phase. Its drawback is the difficulty in evaporation of the butanol component in the organic phase.

3.2. Lipid Separation by HPLC

For separating lipids before mass analysis, high-performance liquid chromatography (HPLC or LC) is a common choice. There are several means of separation, including Hydrophilic Interaction Liquid Chromatography, supercritical fluid chromatography or two-dimensional liquid chromatography. Each of these has shown to be suited for particular uses: NP-HPLC is most effective to separate glycerophospholipids using headgroup polarity, while RP-HPLC uses chain length, degree of unsaturation and substitution to sort out fatty acids like Eicosanoids. The combination of RP and NP or Hydrophilic Interaction Liquid Chromatography (HILC) columns provides wider lipidome coverage for untargeted studies. For general lipodomics studies, nano-flow liquid chromatography (nLC) works to improve lipidome coverage and, in general, the sensitivity of measurements. When the eluate is integrated with the ionization source of a mass spectrometer, online or offline chromatographic (HPLC/UHPLC) separation of lipids is a good choice [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45].
  • Reversed-phase LC. This type of LC works by separating lipids based on their hydrophobicity. Although retention times vary between individual lipids, this remains the most commonly used method for working with complex lipids. Lipidomics research generally employs short columns (range of 50–150 mm; typically, 100); particle sizes either less than 2 µm or fused-core between 2.6 and 2.8 µm; and sorbents modified with C18 or C8
  • Normal-phase LC. This type of LC is a good complement to reversed-phase LC. It uses highly non-polar solvents with low ionization capacity to separate lipids based on their polarities, which makes it an excellent separation mechanism for phospholipids, especially PA. With its longer retention times, generally between 30 and 60 min, it improves separation of large particle sizes on long columns. Lipid classes may be separated using this technique.
  • Hydrophilic interaction liquid chromatography. This technique is more reproducible and robust, with the added benefit of greater compatibility with MS. However, the physical properties of either reversed-phase or normal-phase columns are somewhat compromised.
  • Supercritical fluid chromatography. Using columns packed with sub-2 µm particles, this method is a newly emerging technique that can be used for fast lipid profiling (<20 min).
  • Two-dimensional liquid chromatography. In this technique, the injected sample is separated by passing through two different separation stages. This is accomplished through sequentially connecting two different chromatographic columns, with the effluent from the first system being transferred onto the second column. It can optimize the separation conditions of complex lipidomes. One drawback is that it requires more labor and time.

3.3. Summary of Modern MS-Based Lipidomics Approaches (Table 3)

  • LC-MS/MS
Resolves lipid classes and/or molecular species before MS analysis.
Uses pre-separation of lipids.
Powerful for targeted analysis of very low abundance lipid classes, but slow for comprehensive analysis of cellular lipidomes.
  • Shotgun lipidomics
Uses lipid chemistry and physics for analysis under constant concentrations.
Fairly high throughput, but dynamic range might be affected with coexistence of other major lipids.
  • MALDI-MS
Provides rapid screening and can identify lipid classes with high sensitivity.
Comprehensive analysis is limited.
  • MS imaging
Useful for determining spatial distribution of lipids.
Largely qualitative comparison.
  • Ion-mobility MS
A post-ionization separation technique based on the mass, charge, size and shape of lipids.
Enables class separation analysis of isomeric and isobaric species.
Mass Spectrometry Lipid Analysis.
There are three major categories of MS techniques, each with a different purpose:
  • Global Lipidomic Analysis: This type of approach, referred to as a “shotgun” technique, offers the capability to identify and quantify large numbers of cellular lipids on a high-throughput basis. Several of these techniques are widely used for analyses of wide-ranging pathways and networks involved with metabolism, trafficking and homeostasis of lipids. Recent developments in mapping techniques provide support for research into spatial and temporal relationships among lipids.
  • These methods use LC-MS and LC-MS/MS to identify a Targeted Lipidomic Analysis: much smaller set of lipids of interest.
  • Novel Lipid Discovery: Techniques using LC coupled with MS work with different enrichment technologies with the aim of finding novel classes and molecular species of lipids. This type of analysis involving lipid mediators in biological samples is complicated by their low concentrations in biological samples, their transient nature and limited half-lives.
Targeted versus untargeted approaches. Untargeted analyses are preferred for producing the widest possible scan of lipid ions, useful for examining the lipid composition for discovery efforts. However, these studies are hindered by the limitations of TOF and Orbitrap analyzers, including lower sensitivity and quantitative abilities. Due to this issue, targeted approaches, which involve a predefined list of lipid species of interest, are often used instead. This difficulty is overcome through the use of quadrupole instruments, which apply a specific sequence of voltages, ranging from a few to several thousands, on the mass filters correlated to the lipids of interest. Such ‘multiple reaction monitoring’ allows for more sensitive and accurate quantification of each individual species.
Table 3. MS-based lipidomics approaches.
Table 3. MS-based lipidomics approaches.
ParameterType of MS
Shotgun MSLC-MS/MSMALDI-MSMS ImagingIon-Mobility MS
ReproducibilityHighHighMediumMediumHigh
SensitivityLowHighMediumMediumHigh
ResolutionLowHighHighHighHigh
QuantitativityYesYes, with ISYesYesYes, with IS
Sample preparationMinimalExtensive sample preparation stepsMinimalMinimalExtensive sample preparation steps
Structural informationYesYesNoNoYes
High throughputYes, but dynamic range might be affected with coexistence of other major lipidsYesNoNoYes
Information/Key point - Uses lipid chemistry and physics for analysis under constant concentrations
- Infuse sample directly into the MS
- Less time-consuming and low cost
- High reproducibility
- Low sensitivity
incapable of
distinguishing isomers
- Application-typical building blocks include glycerol, sphingoid bases, polar head groups and fatty acyl substituents (or other aliphatic chains).
- Uses pre-separation of lipids
- Resolves lipid classes and/or molecular species before MS analysis
- Powerful for targeted analysis of very low abundance lipid class, but fairly slow for
comprehensive analysis of cellular lipidomes
- The most commonly used method in lipidomics; able
to analyze a wide variety of lipids
- High separation efficiency
and sensitivity
- Organic solvent
Consumption
Metabolite target analysis
- Metabolite profiling
Lipidomics
- Provides rapid screening and can identify lipid classes with high sensitivity
- Comprehensive analysis is limited
- Useful for determining spatial distribution of lipids
- Largely qualitative comparison
- Ionize sample by coating with matrix and irradiating
laser
- High resolution; the most established technique
- Requires matrix
pretreatment
- A post-ionization separation technique based on the masses, charges, sizes and
shapes of lipids
- Might enable the analysis of isomeric and isobaric species
- Comprehensive analysis under development
LC–MS, liquid chromatography–mass spectrometry; IS, internal standard. MALDI, matrix-assisted laser desorption/ionization; MS, mass spectrometry, HPLC, high-performance liquid chromatography; MSI, mass spectrometry imaging.

3.4. Ionization Technique

  • ESI MS. This technique works with the gaseous ions generated from polar, thermally labile and mainly non-volatile molecules and is useful for a variety of lipids. As a soft-ionization method, its effects on the sample chemistry are minimal prior to analysis. There are a number of ESI-MS techniques which enable analysis of a variety of classes, subclasses and individual lipid species within biological extracts. One technique, referred to as “shotgun lipodomics”, has been developed by Han and his team. For this method, a crude lipid extract is infused directly into an ESI source which is tuned for separation of lipids using their intrinsic electrical properties.
  • APCI MS. While it is similar to ESI, this method relies on the generation of ions as the heated analyte solvent interacts with a corona discharge needle set for a high electrical potential. As primary ions form in the immediate vicinity of the needle, their interaction with the solvent forms secondary ions, which, in turn, ionize the sample. This is a very good technique for the analysis of nonpolar lipids such as triacylglycerols, sterols and fatty acid esters.
  • DESI MS. DESI mass spectrometry is an ambient ionization technique which combines desorption ionization with ESI. This is accomplished by directing an electrically charged mist to the sample surface that is a few millimeters away, making it an excellent tool for mapping lipid distribution contained in biological samples. A major advantage is that, because no matrix is required for tissue preparation, DESI MS allows for multiple consecutive measurements on the same tissue specimen.
  • MALDI MS. This technique, typically employed in the analysis of large proteins, has proven beneficial in lipid studies. A mixture containing a matrix such as 2,5-dihydroxybenzoic acid and the lipid sample is placed on the sample holder in a small spot. When a laser is fired at the spot, the matrix absorbs the energy, subsequently transferring it to the analyte, resulting in ionization of the molecule. MALDI–time-of-flight (MALDI-TOF) MS has shown great promise as an approach for lipidomics studies, particularly for the imaging of lipids from tissue slides.
Spatial lipidomics. With its high sensitivity in the lipid range, DESI is recognized for its strength in the detection and mapping of lipids contained in tissue samples. Additionally, MALDI techniques facilitate direct detection of lipids. When thin tissue slices are coated with such MALDI matrixes, the result is an abundance of lipid-related ions over a sequence of spectra. Detection of phospholipids, sphingolipids and glycerolipids is enhanced in tissues such as heart, kidney and brain.
The majority of lipidomics measures the relative lipid concentration in an extract of a biological sample such as a human tissue. This analysis generates the average lipid profile of the tissue, not considering regional heterogeneity due to cell-type composition. Spatial lipodomics is required to indicate this heterogeneity. The typical set-up for MALDI analysis involves a thin frozen section of a tissue which is fixed to a slide and covered by a thin layer of matrix. When the slide is then scanned by a laser, lipid ions are released. The separation is measured by the time intervals for the individual ions to travel through a vacuum tube; these intervals are dependent on their mass. This method allows the lipid composition of the tissue to be determined pixel by pixel and visualized in 2D as a colored image of the tissue. The latest high-end instruments enable identification of thousands of lipids at a spatial resolution close to a single cell level (10 μm). In addition to plasma, lipidomic profiling of urine and extracellular vesicles is a newer area of interest which may also prove to be informative as a non-invasive source of cancer biomarkers. Table 4 shows lipidomic studies on biofluids reporting diagnostic lipid signatures [46].

4. The Importance of Lipid Metabolism in Oncology

Lipid metabolic reprogramming produces changes in lipid metabolism, which disrupt a variety of cell functions such as cell cycles, proliferation, growth and differentiation; such alterations lead to carcinogenesis. These changes play a significant role in the interaction between the cancer cells and their surrounding microenvironment. There are several types of lipids which, in addition to serving as energy sources and structural components of biological membranes, act as signaling molecules or secondary messengers. Potential lipid biomarkers in the oncology field are shown in Table 5 [47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96].

4.1. Fatty Acid (FA)

Lipid uptake is accomplished through membrane-associated transport proteins such as fatty acid transport protein-1 (FATP1), fatty acid translocase (CD36) and liver fatty acid-binding protein (L-FABP). With regard to FA metabolism, their expression is largely regulated by the master transcription factor of lipogenesis, known as sterol regulatory element-binding protein 1 (SREBP-1). Ovarian cancer, gastric cancer, glioblastoma and oral squamous cell carcinoma all involve upregulation of CD-36 expression. While two early drug candidates targeting this process were unsuccessful due to ineffectiveness and severe side effects, newer formulations are currently undergoing preclinical trials. Changes in enzymes involved in FAO (fatty acid oxidation) are common in cancer cells and ultimately affect this major bioenergetic pathway that promotes proliferation, metastasis, stemness and resistance to treatment for a variety of cancers.
Breast cancer, gastric cancer and prostate cancer have been linked with overexpression of carnitine palmitoyl transferase I (CPT1). It was hoped that CPT1 inhibitors such as etomoxir and phexiline would display antitumor activity; severe adverse reactions have derailed further consideration of them. It has been reported that, in patients with nonalcoholic steatohepatitis (NASH), lipotoxicity for hepatocellular carcinoma (HCC) cells is thwarted by CPT2 downregulation-mediated suppression of FAO. Elevated hepatocarcinogenesis, resulting from accumulation of acylcarnitine as an oncometabolite, as well as a reduction in CPT2 in HCC, have been confirmed. There is a negative correlation between CPT2 expression and tumor stage, while low expression of CPT2 is associated with poor prognosis in colon cancer. There are indications that Th17 inflammation in Type 2 diabetes may be activated by FAO. This is achieved through blocking CPT1A or inhibiting FAO, which reduces the production of IL-17 by T helper 17 (Th17) cells.
Because FAO is implicated in the synthesis of many essential fatty acid derivatives, it is considered to play a vital role in cells. For example, fatty acid is the basic material for prostaglandin and leukotriene synthesis. Fatty acid is part of a hot topic of discussion as to whether FAO is required for regulatory T (Treg) cell differentiation. De novo fatty acid synthesis in CD4+ T cells depends on monounsaturated fatty acids (MUFAs) for endogenous fatty acid synthesis, with stearoyl-CoA desaturases (SCDs) acting as the catalyst. Strong expression of stearoyl-CoA desaturases (SCDs) is shown in human and murine cells, and mouse models have shown that the inhibition of SCDs leads to Tfh cell apoptosis. Additional mouse experiments indicate that phosopholipid production for cell membranes is facilitated by ACC1-mediated de novo fatty acid synthesis and produces Th17 cell differentiation. Th17 immune response, proliferation and infiltration are suppressed through elimination of ACC1 in murine models of colitis. Further studies reveal correlations between high-fat diets and Th17 cell differentiation; ACC1 similarly affects Th17 cells through added expression of IL17A, IL23R, LTB4R1 and CCR6. There is an abundance of reports in the literature stating that naïve T cell activation and differentiation are dependent on the synergistic relationship between the generation of Acetyl–CoA during glycolysis and de novo fatty acid synthesis.

4.2. Cholesterol

Cancer cells require high levels of cholesterol to maintain their rapid proliferation; indeed, increased cholesterol biosynthesis is a hallmark of many cancers. Upregulation of SREBP2 and its target genes is prevalent in prostate cancer, breast cancer and HCC. Activation of the Hedgehog and Notch signaling pathways allows for cholesterol biosynthesis, which, in turn, facilitates cancer stem cell maintenance. Accumulation of cholesteryl esters (CEs) is another characteristic of cancer; for example, Acetyl-CoA acetyltransferase 1 (ACAT1) is correlated with pancreatic cancer and lymphocytic leukemia. Other common components of the tumor microenvironment (TME) include oxysterols; one example is the accumulation of 27-hydroxycholesterol (27HC) in the breast and tumor tissue of estrogen receptor-positive breast cancer patients. One possible anti-tumor therapy involves the inhibition of cholesterol metabolism. Statins, regardless of use before or after diagnosis, are well-known HMGCR inhibitors which reduce mortality in various cancer types and are under clinical investigation as anti-tumor treatments
HMG-CoA reductase, which is essential for de novo cholesterol synthesis in CD4+ T cells, produces mevalonic acid (MVA) from HMG-CoA. The next step along the mevalonate pathway is conversion of MVA to squalene, which is then converted to cholesterol. Acting as an MG-CoA reductase inhibitor by ameliorating hyperlipidemia, statin produces a greater share of Treg cells while suppressing Treg cell formation. With rheumatoid arthritis RA, when Th1 cells are treated with statins, activation of the cholesterol biosynthesis pathway causes them to switch from IFN-γ to IL-10+. This is an indicator that the proinflammatory function of Th1 cells depends on cholesterol biosynthesis.

4.3. Triacylglycerol and Lipid Droplets

Triacylglycerol, or TAG, contains three fatty acids which are esterified to a glycerol molecule; its functions are fatty acid storage and aiding in efficient transport. The rapid proliferation during carcinogenesis increases demand for FA. The production of FA is accomplished through increases in de novo synthesis and uptake from extracellular sources. Autophagy is another source, using lipid droplets to aid in the recycling of intracellular lipids. In addition, lipid droplets are important in defending against reactive oxygen species (ROS) and serve to defend against stress in difficult environments such as hypoxia and low nutrition. There is a correlation between various cancers and excessive concentration of lipid droplets; such accumulations are indicators of a negative prognosis for breast and pancreatic cancers. This shows the value of determining changes in their metabolism. Brain, breast, renal and prostate cancers are all associated with the production of hypoxia-associated lipid droplets. Direct stimulation of Lipin-1 in an HIF-1α-dependent manner results from hypoxia exposure, leading to accumulation of lipid droplets in HCC and cervical cancer. Hypoxia also causes upregulation of FA uptake via FABP3/7 and accumulation of lipid droplets, leading to estrogen receptor (ER) and redox homeostasis during oxygen deprivation. The FA produced during this process is used for mitochondrial energy synthesis, along with increased cell proliferation upon reoxygenation.

4.4. Phospholipid

Additional cell membrane components include phospholipid metabolism phospholipids (PLs), which are diversified in both structure and function. They are involved in homeostasis, cell adhesion and migration, signal transduction, vesicle transport, apoptosis and posttranslational modifications. Subclasses of PLs include phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylserine (PS), phosphatidylglycerol (PG), phosphatidylinositol (PI), and phosphatidic acid (PA), depending on the type of polar head group.
Lysophosphatidylcholine acyltransferase (LPCAT) is gaining recognition as another class of enzyme that appears to be affected in many cancers. LPCAT1 upregulation is connected to clear renal cell carcinoma, oral squamous cell carcinoma, hepatoma, esophageal cancers, gastric cancers, breast cancers, colorectal cancers and prostate cancers. Prognosis and survival for clear cell renal cell carcinoma, breast cancer and prostate cancer are linked with LPCAT1 expression; it is useful as a diagnostic biomarker for prostate and esophageal cancers. Elevated levels of cell proliferation, migration and metastasis in clear cell renal cell carcinoma and HCC cell lines all result from overexpression of LPCAT1. Increased levels of saturated phospholipids are present in clear cell renal cell carcinoma, HCC and gastric cancers, which is aligned with LPCAT1 enzyme activity. Currently the relationship between the behavior of cancer cells and these types of changes in PC levels is now well understood and will require further research.
It is possible that tumor growth may be promoted when LPCAT1 undergoes lyso-PAF acetyltransferase activity to produce PAF, which is important to cell growth as a lipid mediator. Overexpression of LPCAT2 is seen in cervical and breast cancers and is thought to be a susceptibility gene for aggressive prostate cancer in animal models and genome-wide association studies in human patients. This indicates that by targeting LPCAT2-mediated intracellular LD formation, it may be possible to restore sensitivity to chemotherapy in colorectal cancer. When LPCAT3 is reduced in the mouse gut, it produces a reduction in the composition of polyunsaturated phospholipids and promotes tumor development and growth in ApcMin mice. LPCAT4 is associated with intestinal stem cell proliferation and tumorigenesis and is also associated with high levels of PC (16:0/16:1) in colorectal cancer.

4.5. Glycosphingolipids (GSLs)

GSLs are sphingophospholipids which contain glycosyl groups. Gangliosides are sialylated GSLs on the cell membrane of T cells which interact with pathogen antigens. Shigella, the enteroinvasive bacterium responsible for acute rectocolitis, binds to the CD4+ T cell via gangliosides; T cell activation is elevated in the infection mouse model as a result. HIV-1 cell infection is reduced when the ganglioside GM-1 restricts entry. Thus, bacterial and viral infection may be impaired through regulation of GSL metabolism.

5. Lipid Metabolism in Cancer Cells

There are a number of tumor-infiltrating immune cells, or TIIs, which are another component of tumors. Among them are T and B lymphocytes, tumor-associated macrophages (TAMs), dendritic cells (DCs), myeloid-derived suppressor cells (MDSCs), neutrophils and natural killer (NK) cells. Their functions vary from anti-tumor to tumor promoting depending on tumor stage and type. Changes in FA metabolism can alter the polarity of TAMs to a pro-tumoral phenotype. When tumor cells secrete macrophage colony-stimulating factor (M-CSF), FASN upregulation occurs. The result is activation of PPARβ/δ and expression of the anti-inflammatory cytokine IL-10.

6. Lipid Drug Candidates for Cancer

Table 6 shows agents as modulators of lipid metabolism in cancer [98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122]. Given that FAs are key in proliferation and progression of cancer cells, research into therapeutic strategies seeks drug candidates which act on lipogenic enzymes, transcription factors regulating the intracellular lipid homeostasis, or both. There are five target areas under consideration: (a) lipogenic enzymes (FASN, ACLY, ACC); (b) the exogenous lipid uptake (LXR, CD36, FABP4/5); (c) inflammatory signaling pathways (PTGS2); (d) regulation of intracellular lipid homeostasis (PPARγ, CPT1a, lipin2, HSL, MAGAT, DAGAT); and/or (e) saturated vs. unsaturated FAs. While in vitro preclinical and clinical cancer testing models have indicated the efficacy of these strategies, it must be noted that the varied regulatory mechanisms of lipid metabolism produce challenging side effects. A summary of the major drugs being evaluated is included herein.

7. Challenges for Lipidomics

  • The unique nature, complexity and specificity of lipodomics research make it exciting and challenging. For example, mapping of lipidomes is not possible with present equipment and techniques, which are unable to provide the detailed information necessary to quantify individual lipid molecules present in an organism. The diversity of lipid classes and lipid molecular species complicates efforts for the structural identification of lipids by mass spectrometry and renders it impossible to accommodate all lipid classes with a common method for extraction, chromatography and detection. Several methods need to be applied.
  • LC-MS is gaining favor over “shotgun” or direct infusion methods in lipodomics research, given its enhanced specificity and sensitivity. The greater specificity of MS/MS is invaluable for elucidating additional levels of detail which is crucial for identifying individual acyl chains.
  • One of the challenges related to MS-based lipidomics is the complexity of the sample, with a small mass range between 300 and 900 Da. The classical quadrupoles have a limited mass resolution, and, usually, multiple lipid species with similar masses (referred to as isobaric species) are co-detected, hampering the exact assignment of a mass to a specific individual lipid species. The use of mass spectrometers with higher resolution of mass offers a partial solution to this issue. One example is the time-of-flight (TOF) analyzer, which uses the m/z-dependent acceleration of ions in a flight tube; another is the Orbitrap, where ions oscillate around an inner rod. Use of a Fourier transformation with the measured ion oscillations gives the Orbitrap excellent resolution of much smaller mass differences.
  • The differences in ionization efficiency between species present a clear challenge in lipodomics research, though it is enhanced in some situations by the addition of salts such as ammonium acetate. Prior to extraction, one or multiple internal standards at known concentrations for each lipid class of interest are spiked.
  • Reference standards and relevant internal standards remain limited.
  • With the acquisition of the massive output of data generated, and subsequently requiring processing, the voracious demand for computer power and bioinformation clearly represents another major difficulty in MS-based lipodomics research. Spectral alignment and statistical evaluation of fluctuating signal intensities demands significant effort during the collection of chromatographic and MS data. The causes for these discrepancies include biological variations, sample handling issues and analytical accuracy; reliable resolution of lipid levels within a complex mixture may require additional replications.
  • Software packages offer help for identifying correlations between lipid metabolites and physiological phenotypes, especially for developing lipid-based biomarkers. Given the extreme complexity of some lipid pathways, such as the mammalian glycosphingolipid pathway, information technology in lipodomics should help in constructing metabolic maps from data on lipid structures and lipid-related proteins and genes.
  • There is a need to create searchable and interactive databases of lipids and lipid-related genes/proteins. Such databases can be integrated with MS and other experimental data, as well as with metabolic networks. These combined resources will lead to improved therapeutic strategies to prevent or reverse these pathological states involving dysfunction of lipid-related processes.

8. Common Software and Databases Used in Lipidomics

A list of common software databases and their websites used in lipidomics is shown below: All URLs were accessed on 14 March 2024

9. Conclusions

Herein, we present lipid classification and functions, and lipid extraction and analysis techniques used in lipodomics analysis, including tandem MS, MS/MS or MSn [123,124,125,126]. We shed light on ways in which mass spectrometric technology has become much more efficient in characterizing and identifying molecular lipid species in total lipid extracts. Mass spectrometry benefits from its capability to use the mass-to-charge ratios (m/z) of charged ionized analytes for separation and characterization. Fragmentation of lipid ions by means of collision-induced dissociation (CID) generates structural information. These methods, which offer supreme selectivity, sensitivity and the ability to record these fragmentation reactions, provide structural information for components in complex mixtures. Using a targeted approach, as opposed to a global one, lipid analysis provides another method in lipidomics.
It has been reported that remodeling, reprogramming and metabolism to signaling are some of the pathways for cancer cells to undergo profound alterations in lipid homeostasis. T cell differentiation, functions and survival are increasingly demonstrated to be linked to processes of lipid metabolism. Table 7 shows lipid metabolism in functional subsets of T cells. In this review, we have shown that lipidomics provides a promising and potentially powerful tool for the development of new drugs which are targeted to treat cancers. Also, analyzing the relative lipid content in tumor and normal tissue microenvironments is proving helpful in narrowing our search for more precisely targeted lipid metabolism regulation strategies. In addition to its role as an essential component of cancer cell growth and survival, lipid metabolism is involved in the crosstalk with immune cells in the TME. As we improve our knowledge of mechanisms of lipid metabolism in cancer, we seek to gain a better understanding of related enzymes and genetic factors contributing to their metabolism. This knowledge offers the potential for the development of newer, more powerful and precise treatments to fight cancer and restore tumor immunology.

Author Contributions

Data Curation: A.M. and T.Z. Writing—Original Draft Preparation: A.F.N. Writing—Review and Editing: P.N. and H.O. Visualization: X.N., S.L., M.U.B., J.Y., L.C. and J.H. Supervision: L.C. Funding Acquisition: L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Lung SPORE, NIH, GR120384.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Griffths, W.J.; Ogundare, M.; Williams, C.M.; Wang, Y. On the future of ‘omics’: Lipidomics. J. Inherit. Metab. Dis. 2011, 34, 583–592. [Google Scholar] [CrossRef] [PubMed]
  2. Fahy, E.; Subramaniam, S.; Brown, H.A.; Glass, C.K.; Merrill, A.H.; Murphy, R.C.; Raetz, C.R.; Russell, D.W.; Seyama, Y.; Shaw, W.; et al. A comprehensive classification system for lipids. J. Lipid Res. 2005, 46, 839–861. [Google Scholar] [CrossRef] [PubMed]
  3. Han, X. Lipidomics: Comprehensive Mass Spectrometry of Lipids/Xianlin Han; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2016. [Google Scholar]
  4. Bach D and Wachtel E, Phospholipid/cholesterol model membranes: Formation of cholesterol crystallites. Biochim. Biophys. Acta 2003, 1610, 187–197. [CrossRef] [PubMed]
  5. Han, X.; Gross, R.W. The Foundations and Development of Lipidomics. J. Lipid Res. 2022, 63, 100164. [Google Scholar] [CrossRef] [PubMed]
  6. Tsai MJ and O’Malley BW, Molecular mechanisms of action of steroid/thyroid receptor superfamily members. Annu. Rev. Biochem. 1994, 63, 451–486. [CrossRef] [PubMed]
  7. Kuzuyama, T.; Seto, H. Diversity of the biosynthesis of the isoprene units. Nat. Prod. Rep. 2003, 20, 171–183. [Google Scholar] [CrossRef] [PubMed]
  8. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  9. Beloribi-Djefafia, S.; Vasseur, S.; Guillaumond, F. Lipid metabolic reprogramming in cancer cells. Oncogenesis 2016, 5, e189. [Google Scholar] [CrossRef] [PubMed]
  10. van Meer, G.; Voelker, D.R.; Feigenson, G.W. Membrane lipids: Where they are and how they behave. Nat. Rev. Mol. Cell Biol. 2008, 9, 112–124. [Google Scholar] [CrossRef]
  11. Addepalli, R.V.; Mullangi, R. A concise review on lipidomics analysis in biological samples. ADMET DMPK 2021, 9, 1–22. [Google Scholar]
  12. Sorgi, C.A.; Peti, A.P.; Petta, T.; Meirelles, A.F.; Fontanari, C.; Moraes, L.A.; Faccioli, L.H. Comprehensive high-resolution multiple-reaction monitoring mass spectrometry for targeted eicosanoid assays. Sci. Data 2018, 5, 180167. [Google Scholar] [CrossRef]
  13. Züllig, T.; Trötzmüller, M.; Köfeler, H.C. Lipidomics from sample preparation to data analysis: A primer. Anal. Bioanal. Chem. 2020, 412, 2191–2209. [Google Scholar] [CrossRef]
  14. Seppanen-Laakso, T.; Oresic, M. How to study lipidomes. J. Mol. Endocrinol. 2009, 42, 185–190. [Google Scholar] [CrossRef]
  15. Folch, J.; Lees, M.; Sloane Stanley, G.H. A simple method for the isolation and purification of total lipides from animal tissues. J. Biol. Chem. 1957, 226, 497–509. [Google Scholar] [CrossRef]
  16. Bligh, E.G.; Dyer, W.J. A rapid method of total lipid extraction and purification. J. Biochem. Physiol. 1959, 37, 911–917. [Google Scholar]
  17. Del Boccio, P.; Pieragostino, D.; di Ioia, M.; Petrucci, F.; Lugaresi, A.; De Luca, G.; Gambi, D.; Onofrj, M.; Di Ilio, C.; Sacchetta, P.; et al. Lipidomic investigations for the characterization of circulating serum lipids in multiple sclerosis. J. Proteom. 2011, 74, 2826–2836. [Google Scholar] [CrossRef]
  18. Yang, K.; Han, X. Lipidomics: Techniques, applications, and outcomes related to biomedical sciences. Trends Biochem. Sci. 2016, 41, 954–969. [Google Scholar] [CrossRef]
  19. Wang, M.; Han, X. Multidimensional mass spectrometry-based shotgun lipidomics. Methods Mol. Biol. 2014, 1198, 203–220. [Google Scholar]
  20. Blanksby, S.J.; Mitchell, T.W. Advances in Mass Spectrometry for Lipidomics. Annu. Rev. Anal. Chem. 2010, 3, 433–465. [Google Scholar] [CrossRef]
  21. Harkewicz, R.; Dennis, E.A. Applications of Mass Spectrometry to Lipids and Membranes. Annu. Rev. Biochem. 2011, 80, 301–325. [Google Scholar]
  22. Dennis, E.A.; Norris, P.C. Eicosanoid storm in infection and inflammation. Nat. Rev. Immunol. 2015, 15, 511–523. [Google Scholar] [CrossRef]
  23. Sandra, K.; Sandra, P. Lipidomics from an analytical perspective. Curr. Opin. Chem. Biol. 2013, 17, 847–853. [Google Scholar] [CrossRef]
  24. Massey, K.A.; Nicolaou, A. Lipidomics of polyunsaturated-fatty-acid-derived oxygenated metabolites. Biochem. Soc. Trans. 2011, 39, 1240–1246. [Google Scholar] [CrossRef]
  25. Sun, T.; Chen, J.; Yang, F.; Zhang, G.; Chen, J.; Wang, X.; Zhang, J. Lipidomics reveals new lipid-based lung adenocarcinoma early diagnosis model. EMBO Mol. Med. 2024, 16, 854–869. [Google Scholar] [CrossRef]
  26. Wang, C.; Wang, M.; Han, X. Applications of mass spectrometry for cellular lipid analysis. Mol. BioSyst. 2015, 11, 698–713. [Google Scholar] [CrossRef]
  27. Pulfer, M.; Murphy, R.C. Electrospray mass spectrometry of phospholipids. Mass Spectrom. Rev. 2003, 22, 332–364. [Google Scholar] [CrossRef]
  28. Fuchs, B.; Suss, R.; Schiller, J. An update of MALDI-TOF mass spectrometry in lipid research. Prog. Lipid Res. 2011, 49, 450–475. [Google Scholar] [CrossRef]
  29. Serna, J.; Garcia-Seisdedos, D.; Alcazar, A.; Lasuncion, M.A.; Busto, R.; Pastor, O. Quantitative lipidomic analysis of plasma and plasma lipoproteins using MALDI-TOF mass spectrometry. Chem. Phys. Lipids 2015, 189, 7–18. [Google Scholar] [CrossRef]
  30. Han, X.; Yang, K.; Gross, R.W. Multi-dimensional Mass Spectrometry-based Shotgun Lipidomics and Novel Strategies for Lipidomic Analyses. Mass Spectrom. Rev. 2012, 31, 134–178. [Google Scholar] [CrossRef]
  31. Wang, M.; Wang, C.; Han, R.H.; Han, X. Novel Advances in Shotgun Lipidomics for Biology and Medicine. Prog. Lipid Res. 2016, 61, 83–108. [Google Scholar] [CrossRef]
  32. Hu, T.; Zhang, J. Mass-spectrometry-based Lipidomics. J. Sep. Sci. 2018, 41, 351–372. [Google Scholar] [CrossRef]
  33. Eberlin, L.S.; Dill, A.L.; Costa, A.B.; Ifa, D.R.; Cheng, L.; Masterson, T.; Koch, M.; Ratliff, T.L.; Cooks, R.G. Cholesterol Sulfate Imaging in Human Prostate Cancer Tissue by Desorption Electrospray Ionization Mass Spectrometry. Anal. Chem. 2010, 82, 3430–3434. [Google Scholar] [CrossRef]
  34. Angerer, T.B.; Magnusson, Y.; Landberg, G.; Fletcher, J.S. Lipid Heterogeneity Resulting from Fatty Acid Processing in the Human Breast Cancer Microenvironment Identified by GCIB-ToF-SIMS Imaging. Anal. Chem. 2016, 88, 11946–11954. [Google Scholar] [CrossRef]
  35. Banerjee, S.; Zare, R.N.; Tibshirani, R.J.; Kunder, C.A.; Nolley, R.; Fan, R.; Brooks, J.D.; Sonn, G.A. Diagnosis of Prostate Cancer by Desorption Electrospray Ionization Mass Spectrometric Imaging of Small Metabolites and Lipids. Proc. Natl. Acad. Sci. USA 2017, 114, 3334–3339. [Google Scholar] [CrossRef]
  36. Ling, Y.S.; Liang, H.; Lin, M.; Tang, C.; Wu, K.; Kuo, M.; Lin, C.Y. Two-dimensional LC-MS/MS to Enhance Ceramide and Phosphatidylcholine Species Profiling in Mouse Liver. Biomed. Chromatogr. 2014, 28, 1284–1293. [Google Scholar] [CrossRef]
  37. Sarafian, M.H.; Gaudin, M.; Lewis, M.R.; Martin, F.P.; Holmes, E.; Nicholson, J.K.; Dumas, M.E. Objective Set of Criteria for Optimization of Sample Preparation Procedures for Ultra-High Throughput Untargeted Blood Plasma Lipid Profiling by Ultra Performance Liquid Chromatography−Mass Spectrometry. Anal. Chem. 2014, 86, 5766–5774. [Google Scholar] [CrossRef]
  38. Nassar, A.-E.F.; Talaat, R. Strategies for Dealing with Metabolite Elucidation in Drug Discovery and Development. Drug Discov. Today 2004, 9, 317–327. [Google Scholar] [CrossRef]
  39. Shimma, S.; Sugiura, Y.; Hayasaka, T.; Hoshikawa, Y.; Noda, T.; Setou, M. MALDI-based imaging mass spectrometry revealed abnormal distribution of phospholipids in colon cancer liver metastasis. J. Chromatogr. 2007, B855, 98–103. [Google Scholar] [CrossRef]
  40. Kendall, A.C.; Koszyczarek, M.M.; Jones, E.A.; Hart, P.J.; Towers, M.; Griffiths, C.E.; Morris, M.; Nicolaou, A. Lipidomics for translational skin research: A primer for the uninitiated. Exp. Dermatol. 2018, 27, 721–728. [Google Scholar] [CrossRef]
  41. Goto-Inoue, N.; Hayasaka, T.; Zaima, N.; Setou, M. Imaging mass spectrometry for lipidomics. Biochim. Et Biophys. Acta –Mol. Cell Biol. Lipids 2011, 1811, 961–969. [Google Scholar] [CrossRef]
  42. Bokhart, M.T.; Muddiman, D.C. Infrared matrix-assisted laser desorption electrospray ionization mass spectrometry imaging analysis of biospecimens. Analyst 2016, 141, 5236–5245. [Google Scholar] [CrossRef]
  43. Meier, F.; Garrard, K.P.; Muddiman, D.C. Silver dopants for targeted and untargeted direct analysis of unsaturated lipids via infrared matrix-assisted laser desorption electrospray ionization (IR-MALDESI). Rapid Commun. Mass Spectrom. 2014, 28, 2461–2470. [Google Scholar] [CrossRef]
  44. JonesJones, E.A.; Simon, D.; Karancsi, T.; Balog, J.; Pringle, S.D.; Takats, Z. Matrix assisted rapid evaporation ionisation mass spectrometry. Anal. Chem. 2019, 91, 9784–9791. [Google Scholar] [CrossRef]
  45. Nassar, A.F.; Wu, T.; Nassar, S.F.; Wisnewski, A.V. UPLC–MS for metabolomics: A giant step forward in support of pharmaceutical research. Drug Discov. Today 2016, 22, 463–470. [Google Scholar] [CrossRef]
  46. Butler, L.M.; Perone, Y.; Dehairs, J.; Lupien, L.E.; de Laat, V.; Talebi, A.; Loda, M.; Kinlaw, W.B.; Swinnen, J.V. Lipids and cancer: Emerging roles in pathogenesis, diagnosis and therapeutic intervention. Adv. Drug Deliv. Rev. 2020, 159, 245–293. [Google Scholar]
  47. Hall, Z.; Ament, Z.; Wilson, C.H.; Burkhart, D.L.; Ashmore, T.; Koulman, A.; Littlewood, T.; Evan, G.I.; Griffin, J.L. Myc Expression drives aberrant lipid metabolism in lung cancer. Cancer Res. 2016, 76, 4608–4618. [Google Scholar] [CrossRef]
  48. Lee, G.K.; Lee, H.S.; Park, Y.S.; Lee, J.H.; Lee, S.C.; Lee, J.H.; Lee, S.J.; Shanta, S.R.; Park, H.M.; Kim, H.R.; et al. Lipid MALDI profile classifies non-small cell lung cancers according to the histologic type. Lung Cancer 2012, 76, 197–203. [Google Scholar] [CrossRef]
  49. Pirman, D.A.; Efuet, E.; Ding, X.P.; Pan, Y.; Tan, L.; Fischer, S.M.; DuBois, R.N.; Yang, P. Changes in cancer cell metabolism revealed by direct sample analysis with MALDI mass spectrometry. PLoS ONE 2013, 8, e61379. [Google Scholar] [CrossRef]
  50. Guo, Y.; Wang, X.; Qiu, L.; Qin, X.; Liu, H.; Wang, Y.; Li, F.; Wang, X.; Chen, G.; Song, G.; et al. Probing gender-specific lipid metabolites and diagnostic biomarkers for lung cancer using Fourier transform ion cyclotron resonance mass spectrometry. Clin. Chim. Acta 2012, 414, 135–141. [Google Scholar] [CrossRef]
  51. Marien, E.; Meister, M.; Muley, T.; Fieuws, S.; Bordel, S.; Derua, R.; Spraggins, J.; van de Plas, R.; Dehairs, J.; Wouters, J.; et al. Non-small cell lung cancer is characterized by dramatic changes in phospholipid profiles. Int. J. Cancer 2015, 137, 1539–1548. [Google Scholar] [CrossRef]
  52. Hilvo, M.; Denkert, C.; Lehtinen, L.; Muller, B.; Brockmoller, S.; Seppanen-Laakso, T.; Budczies, J.; Bucher, E.; Yetukuri, L.; Castillo, S.; et al. Novel theranostic opportunities offered by characterization of altered membrane lipid metabolism in breast cancer progression. Cancer Res. 2011, 71, 3236–3245. [Google Scholar] [CrossRef]
  53. Kang, H.S.; Lee, S.C.; Park, Y.S.; Jeon, Y.E.; Lee, J.H.; Jung, S.Y.; Park, I.H.; Jang, S.H.; Park, H.M.; Yoo, C.W.; et al. Protein and lipid MALDI profiles classify breast cancers according to the intrinsic subtype. BMC Cancer 2011, 11, 465–474. [Google Scholar] [CrossRef]
  54. Chughtai, K.; Jiang, L.; Greenwood, T.R.; Glunde, K.; Heeren, R.M. Mass spectrometry images acylcarnitines, phosphatidylcholines, and sphingomyelin in MDA-MB-231 breast tumor models. J. Lipid Res. 2013, 54, 333–344. [Google Scholar] [CrossRef]
  55. Min, H.K.; Kong, G.; Moon, M.H. Quantitative analysis of urinary phospholipids found in patients with breast cancer by nanoflow liquid chromatography—tandem mass spectrometry: II. Negative ion mode analysis of four phospholipid classes. Anal. Bioanal. Chem. 2010, 396, 1273–1280. [Google Scholar] [CrossRef]
  56. Kawashima, M.; Iwamoto, N.; Kawaguchi-Sakita, N.; Sugimoto, M.; Ueno, T.; Mikami, Y.; Terasawa, K.; Sato, T.A.; Tanaka, K.; Shimizu, K.; et al. High-resolution imaging mass spectrometry reveals detailed spatial distribution of phosphatidylinositols in human breast cancer. Cancer Sci. 2013, 104, 1372–1379. [Google Scholar] [CrossRef]
  57. Doria, M.L.; Cotrim, C.Z.; Simoes, C.; Macedo, B.; Domingues, P.; Domingues, M.R.; Helguero, L.A. Lipidomic analysis of phospholipids from human mammary epithelial and breast cancer cell lines. J. Cell. Physiol. 2013, 228, 457–468. [Google Scholar] [CrossRef]
  58. Kim, H.Y.; Lee, K.M.; Kim, S.H.; Kwon, Y.J.; Chun, Y.J.; Choi, H.K. Comparative metabolic and lipidomic profiling of human breast cancer cells with different metastatic potentials. Oncotarget 2016, 7, 67111–67128. [Google Scholar] [CrossRef]
  59. Cifkova, E.; Holcapek, M.; Lisa, M.; Vrana, D.; Gatek, J.; Melichar, B. Determination of lipidomic differences between human breast cancer and surrounding normal tissues using HILIC-HPLC/ESI-MS and multivariate data analysis. Anal. Bioanal. Chem. 2015, 407, 991–1002. [Google Scholar] [CrossRef]
  60. Hilvo, M.; Gade, S.; Hyotylainen, T.; Nekljudova, V.; Seppanen-Laakso, T.; Sysi-Aho, M.; Untch, M.; Huober, J.; von Minckwitz, G.; Denkert, C.; et al. Monounsaturated fatty acids in serum triacylglycerols are associated with response to neoadjuvant chemotherapy in breast cancer patients. Int. J. Cancer 2014, 134, 1725–1733. [Google Scholar] [CrossRef]
  61. Wei, S.; Liu, L.; Zhang, J.; Bowers, J.; Gowda, G.A.; Seeger, H.; Fehm, T.; Neubauer, H.J.; Vogel, U.; Clare, S.E.; et al. Metabolomics approach for predicting response to neoadjuvant chemotherapy for breast cancer. Mol. Oncol. 2013, 7, 297–307. [Google Scholar] [CrossRef]
  62. Min, H.K.; Lim, S.; Chung, B.C.; Moon, M.H. Shotgun lipidomics for candidate biomarkers of urinary phospholipids in prostate cancer. Anal. Bioanal. Chem. 2011, 399, 823–830. [Google Scholar] [CrossRef]
  63. Goto, T.; Terada, N.; Inoue, T.; Nakayama, K.; Okada, Y.; Yoshikawa, T.; Miyazaki, Y.; Uegaki, M.; Sumiyoshi, S.; Kobayashi, T.; et al. The expression profile of phosphatidylinositol in high spatial resolution imaging mass spectrometry as a potential biomarker for prostate cancer. PLoS ONE 2014, 9, e90242. [Google Scholar] [CrossRef]
  64. Goto, T.; Terada, N.; Inoue, T.; Kobayashi, T.; Nakayama, K.; Okada, Y.; Yoshikawa, T.; Miyazaki, Y.; Uegaki, M.; Utsunomiya, N.; et al. Decreased expression of lysophosphatidylcholine (16:0/OH) in high resolution imaging mass spectrometry independently predicts biochemical recurrence after surgical treatment for prostate cancer. Prostate 2015, 75, 1821–1830. [Google Scholar] [CrossRef]
  65. Patel, N.; Vogel, R.; Chandra-Kuntal, K.; Glasgow, W.; Kelavkar, U. A novel three serum phospholipid panel differentiates normal individuals from those with prostate cancer. PLoS ONE 2014, 9, e88841. [Google Scholar] [CrossRef]
  66. Duscharla, D.; Bhumireddy, S.R.; Lakshetti, S.; Pospisil, H.; Murthy, P.V.; Walther, R.; Sripadi, P.; Ummanni, R. Prostate cancer associated lipid signatures in serum studied by ESI-tandem mass spectrometryas potential new biomarkers. PLoS ONE 2016, 11, e0150253. [Google Scholar] [CrossRef]
  67. Zhao, Z.; Xiao, Y.; Elson, P.; Tan, H.; Plummer, S.J.; Berk, M.; Aung, P.P.; Lavery, I.C.; Achkar, J.P.; Li, L.; et al. Plasma lysophosphatidylcholine levels: Potential biomarkers for colorectal cancer. J. Clin. Oncol. 2007, 25, 2696–2701. [Google Scholar] [CrossRef]
  68. Dobrzynska, I.; Szachowicz-Petelska, B.; Sulkowski, S.; Figaszewski, Z. Changes in electric charge and phospholipids composition in human colorectal cancer cells. Mol. Cell. Biochem. 2005, 276, 113–119. [Google Scholar] [CrossRef]
  69. Kurabe, N.; Hayasaka, T.; Ogawa, M.; Masaki, N.; Ide, Y.; Waki, M.; Nakamura, T.; Kurachi, K.; Kahyo, T.; Shinmura, K.; et al. Accumulated phosphatidylcholine (16:0/16:1) in human colorectal cancer; possible involvement of LPCAT4. Cancer Sci. 2013, 104, 1295–1302. [Google Scholar] [CrossRef]
  70. Mirnezami, R.; Spagou, K.; Vorkas, P.A.; Lewis, M.R.; Kinross, J.; Want, E.; Shion, H.; Goldin, R.D.; Darzi, A.; Takats, Z.; et al. Chemical mapping of the colorectal cancer microenvironment via MALDI imaging mass spectrometry (MALDI-MSI) reveals novel cancer-associated Feld effects. Mol. Oncol. 2014, 8, 39–49. [Google Scholar] [CrossRef]
  71. Thomas, A.; Patterson, N.H.; Marcinkiewicz, M.M.; Lazaris, A.; Metrakos, P.; Chaurand, P. Histology-driven data mining of lipid signatures from multiple imaging mass spectrometry analyses: Application to human colorectal cancer liver metastasis biopsies. Anal. Chem. 2013, 85, 2860–2866. [Google Scholar] [CrossRef]
  72. Coviello, G.; Tutino, V.; Notarnicola, M.; Caruso, M.G. Erythrocyte membrane fatty acids profile in colorectal cancer patients: A preliminary study. Anticancer. Res. 2014, 34, 4775–4779. [Google Scholar]
  73. Zhang, Y.; Liu, Y.; Li, L.; Wei, J.; Xiong, S.; Zhao, Z. High resolution mass spectrometry coupled with multivariate data analysis revealing plasma lipidomic alteration in ovarian cancer in Asian women. Talanta 2016, 150, 88–96. [Google Scholar] [CrossRef]
  74. Zhao, Z.; Cai, Q.; Xu, Y. The Lipidomic analyses in low and highly aggressive ovarian cancer cell lines. Lipids 2016, 51, 179–187. [Google Scholar] [CrossRef]
  75. Kang, S.; Lee, A.; Park, Y.S.; Lee, S.C.; Park, S.Y.; Han, S.Y.; Kim, K.P.; Kim, Y.H.; Yoo, C.W.; Kim, H.K. Alteration in lipid and protein profiles of ovarian cancer: Similarity to breast cancer. Int. J. Gynecol. Cancer 2011, 21, 1566–1572. [Google Scholar] [CrossRef] [PubMed]
  76. Sutphen, R.; Xu, Y.; Wilbanks, G.D.; Fiorica, J.; Grendys, E.C., Jr.; LaPolla, J.P.; Arango, H.; Hoffman, M.S.; Martino, M.; Wakeley, K.; et al. Lysophospholipids are potential biomarkers of ovarian cancer. Cancer Epidemiol. Biomark. Prev. 2004, 13, 1185–1191. [Google Scholar] [CrossRef]
  77. Xu, Y.; Shen, Z.; Wiper, D.W.; Wu, M.; Morton, R.E.; Elson, P.; Kennedy, A.W.; Belinson, J.; Markman, M.; Casey, G. Lysophosphatidic acid as a potential biomarker for ovarian and other gynecologic cancers. JAMA 1998, 280, 719–723. [Google Scholar] [CrossRef]
  78. Xiao, Y.; Chen, Y.; Kennedy, A.W.; Belinson, J.; Xu, Y. Evaluation of plasma lysophospholipids for diagnostic significance using electrospray ionization mass spectrometry (ESI-MS) analyses. Ann. N. Y. Acad. Sci. 2000, 905, 242–259. [Google Scholar] [CrossRef] [PubMed]
  79. Sedlakova, I.; Vavrova, J.; Tosner, J.; Hanousek, L. Lysophosphatidic acid: An ovarian cancer marker. Eur. J. Gynaecol. Oncol. 2008, 29, 511–514. [Google Scholar] [PubMed]
  80. Sedlakova, I.; Vavrova, J.; Tosner, J.; Hanousek, L. Lysophosphatidic acid (LPA)—A perspective marker in ovarian cancer. Tumour Biol. 2011, 32, 311–316. [Google Scholar] [CrossRef] [PubMed]
  81. Zhao, S.; Wang, Y.; Dou, A.; Chen, J.; Lu, X.; Cao, R.; Xu, C.; Xu, G. Study of phospholipid profile of ovarian tumor by high performance liquid chromatography-mass spectrometry. Chin. J. Chromatogr. 2011, 29, 843–850. [Google Scholar]
  82. Jiang, Y.; DiVittore, N.A.; Young, M.M.; Jia, Z.; Xie, K.; Ritty, T.M.; Kester, M.; Fox, T.E. Altered sphingolipid metabolism in patients with metastatic pancreatic cancer. Biomolecules 2013, 3, 435–448. [Google Scholar] [CrossRef] [PubMed]
  83. Liao, Y.; Mu, G.; Zhang, L.; Zhou, W.; Zhang, J.; Yu, H. Lysophosphatidic acid stimulates activation of focal adhesion kinase and paxillin and promotes cell motility, via LPA1–3, in human pancreatic cancer. Dig. Dis. Sci. 2013, 58, 3524–3533. [Google Scholar] [CrossRef] [PubMed]
  84. Macasek, J.; Vecka, M.; Zak, A.; Urbanek, M.; Krechler, T.; Petruzelka, L.; Stankova, B.; Zeman, M. Plasma fatty acid composition in patients with pancreatic cancer: Correlations to clinical parameters. Nutr. Cancer 2012, 64, 946–955. [Google Scholar] [CrossRef] [PubMed]
  85. Uehara, T.; Kikuchi, H.; Miyazaki, S.; Iino, I.; Setoguchi, T.; Hiramatsu, Y.; Ohta, M.; Kamiya, K.; Morita, Y.; Tanaka, H.; et al. Overexpression of lysophosphatidylcholine acyltransferase 1 and concomitant lipid alterations in gastric cancer. Ann. Surg. Oncol. 2016, 23, 206–213. [Google Scholar] [CrossRef] [PubMed]
  86. Dill, A.L.; Eberlin, L.S.; Costa, A.B.; Zheng, C.; Ifa, D.R.; Cheng, L.; Masterson, T.A.; Koch, M.O.; Vitek, O.; Cooks, R.G. Multivariate statistical identification of human bladder carcinomas using ambient ionization imaging mass spectrometry. Chemistry 2011, 17, 2897–2902. [Google Scholar] [CrossRef]
  87. Dill, A.L.; Ifa, D.R.; Manicke, N.E.; Costa, A.B.; Ramos-Vara, J.A.; Knapp, D.W.; Cooks, R.G. Lipid profiles of canine invasive transitional cell carcinoma of the urinary bladder and adjacent normal tissue by desorption electrospray ionization imaging mass spectrometry. Anal. Chem. 2009, 81, 8758–8764. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, J.; Chen, Y.; Zhang, R.; Song, Y.; Cao, J.; Bi, N.; Wang, J.; He, J.; Bai, J.; Dong, L.; et al. Global and targeted metabolomics of esophageal squamous cell carcinoma discovers potential diagnostic and therapeutic biomarkers. Mol. Cell. Proteom. 2013, 12, 1306–1318. [Google Scholar] [CrossRef] [PubMed]
  89. Uchiyama, Y.; Hayasaka, T.; Masaki, N.; Watanabe, Y.; Masumoto, K.; Nagata, T.; Katou, F.; Setou, M. Imaging mass spectrometry distinguished the cancer and stromal regions of oral squamous cell carcinoma by visualizing phosphatidylcholine (16:0/16:1) and phosphatidylcholine (18:1/20:4). Anal. Bioanal. Chem. 2014, 406, 1307–1316. [Google Scholar] [CrossRef] [PubMed]
  90. Moch, H.; Cubilla, A.L.; Humphrey, P.A.; Reuter, V.E.; Ulbright, T.M. The 2016 WHO classification of tumours of the urinary system and male genital organs-part A: Renal, penile, and testicular tumours. Eur. Urol. 2016, 70, 93–105. [Google Scholar] [CrossRef]
  91. Lin, L.; Huang, Z.; Gao, Y.; Chen, Y.; Hang, W.; Xing, J.; Yan, X. LC-MS-based serum metabolic profiling for genitourinary cancer classification and cancer type-specific biomarker discovery. Proteomics 2012, 12, 2238–2246. [Google Scholar] [CrossRef]
  92. Saito, K.; Arai, E.; Maekawa, K.; Ishikawa, M.; Fujimoto, H.; Taguchi, R.; Matsumoto, K.; Kanai, Y.; Saito, Y. Lipidomic signatures and associated transcriptomic profiles of clear cell renal cell carcinoma. Sci. Rep. 2016, 6, 28932–28944. [Google Scholar] [CrossRef] [PubMed]
  93. Cifkova, E.; Holcapek, M.; Lisa, M.; Vrana, D.; Melichar, B.; Student, V. Lipidomic differentiation between human kidney tumors and surrounding normal tissues using HILIC-HPLC/ESI-MS and multivariate data analysis. J. Chromatogr. B 2015, 1000, 14–21. [Google Scholar] [CrossRef] [PubMed]
  94. Dill, A.L.; Eberlin, L.S.; Zheng, C.; Costa, A.B.; Ifa, D.R.; Cheng, L.; Masterson, T.A.; Koch, M.O.; Vitek, O.; Cooks, R.G. Multivariate statistical differentiation of renal cell carcinomas based on lipidomic analysis by ambient ionization imaging mass spectrometry. Anal. Bioanal. Chem. 2010, 398, 2969–2978. [Google Scholar] [CrossRef] [PubMed]
  95. Ishikawa, S.; Tateya, I.; Hayasaka, T.; Masaki, N.; Takizawa, Y.; Ohno, S.; Kojima, T.; Kitani, Y.; Kitamura, M.; Hirano, S.; et al. Increased expression of phosphatidylcholine (16:0/18:1) and (16:0/18:2) in thyroid papillary cancer. PLoS ONE 2012, 7, e48873. [Google Scholar] [CrossRef] [PubMed]
  96. Guo, S.; Qiu, L.; Wang, Y.; Qin, X.; Liu, H.; He, M.; Zhang, Y.; Li, Z.; Chen, X. Tissue imaging and serum lipidomic profiling for screening potential biomarkers of thyroid tumors by matrix-assisted laser desorption/ionization—Fourier transform ion cyclotron resonance mass spectrometry. Anal. Bioanal. Chem. 2014, 406, 4357–4370. [Google Scholar] [CrossRef] [PubMed]
  97. Francesca Perrotti, Consuelo Rosa, Ilaria Cicalini, Paolo Sacchetta, Piero Del Boccio, Domenico Genovesiand Damiana Pieragostino, Advances in Lipidomics for Cancer Biomarkers Discovery. Int. J. Mol. Sci. 2016, 17, 1992.
  98. Schcolnik-Cabrera, A.; Chavez-Blanco, A.; Dominguez-Gomez, G.; Taja-Chayeb, L.; Morales-Barcenas, R.; Trejo-Becerril, C.; Perez-Cardenas, E.; Gonzalez-Fierro, A.; Dueñas-González, A. Orlistat as a FASN inhibitor and multitargeted agent for cancer therapy. Expert. Opin. Investig. Drugs 2018, 2, 475–489. [Google Scholar] [CrossRef] [PubMed]
  99. de Aquino, I.G.; Bastos, D.C.; Cuadra-Zelaya, F.J.; Teixeira, I.F.; Salo, T.; Della Coletta, R.; Graner, E. Anticancer properties of the fatty acid synthase inhibitor TVB-3166 on oral squamous cell carcinoma cell lines. Arch. Oral Biol. 2020, 113, 104707. [Google Scholar] [CrossRef] [PubMed]
  100. Zaytseva, Y.Y.; Rychahou, P.G.; Le, A.T.; Scott, T.L.; Flight, R.M.; Kim, J.T.; Harris, J.; Liu, J.; Wang, C.; Morris, A.J.; et al. Preclinical evaluation of novel fatty acid synthase inhibitors in primary colorectal cancer cells and a patient-derived xenograft model of colorectal cancer. Oncotarget 2018, 9, 24787–24800. [Google Scholar] [CrossRef]
  101. Zadra, G.; Ribeiro, C.F.; Chetta, P.; Ho, Y.; Cacciatore, S.; Gao, X.; Syamala, S.; Bango, C.; Photopoulos, C.; Huang, Y.; et al. Inhibition of de novo lipogenesis targets androgen receptor signaling in castrationresistant prostate cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 631–640. [Google Scholar] [CrossRef]
  102. Wang, C.; Xu, C.; Sun, M.; Luo, D.; Liao, D.F.; Cao, D. Acetyl-CoA carboxylasealpha inhibitor TOFA induces human cancer cell apoptosis. Biochem. Biophys. Res. Commun. 2009, 385, 302–306. [Google Scholar] [CrossRef]
  103. Beckers, A.; Organe, S.; Timmermans, L.; Scheys, K.; Peeters, A.; Brusselmans, K.; Verhoeven, G.; Swinnen, J.V. Chemical inhibition of acetyl-CoA carboxylase induces growth arrest and cytotoxicity selectively in cancer cells. Cancer Res. 2007, 67, 8180–8187. [Google Scholar] [CrossRef]
  104. Svensson, R.U.; Parker, S.J.; Eichner, L.J.; Kolar, M.J.; Wallace, M.; Brun, S.N.; Lombardo, P.S.; Van Nostrand, J.L.; Hutchins, A.; Vera, L.; et al. Inhibition of acetyl-CoA carboxylase suppresses fatty acid synthesis and tumor growth of non-small-cell lung cancer in preclinical models. Nat. Med. 2016, 22, 1108–1119. [Google Scholar] [CrossRef] [PubMed]
  105. Watt, M.J.; Clark, A.K.; Selth, L.A.; Haynes, V.R.; Lister, N.; Rebello, R.; Porter, L.H.; Niranjan, B.; Whitby, S.T.; Lo, J.; et al. Suppressing fatty acid uptake has therapeutic effects in preclinical models of prostate cancer. Sci. Transl. Med. 2019, 11, eaau5758. [Google Scholar] [CrossRef] [PubMed]
  106. Li, X.; Chen, Y.T.; Hu, P.; Huang, W.C. Fatostatin displays high antitumor activity in prostate cancer by blocking SREBP-regulated metabolic pathways and androgen receptor signaling. Mol. Cancer Ther. 2014, 13, 855–866. [Google Scholar] [CrossRef] [PubMed]
  107. Talebi, A.; Dehairs, J.; Rambow, F.; Rogiers, A.; Nittner, D.; Derua, R.; Vanderhoydonc, F.; Duarte, J.A.; Bosisio, F.; Van den Eynde, K.; et al. Sustained SREBP-1-dependent lipogenesis as a key mediator of resistance to BRAF-targeted therapy. Nat. Commun. 2018, 9, 2500. [Google Scholar] [CrossRef] [PubMed]
  108. Hawkins, J.L.; Robbins, M.D.; Warren, L.C.; Xia, D.; Petras, S.F.; Valentine, J.J.; Varghese, A.H.; Wang, K.; Subashi, T.A.; Shelly, L.D.; et al. Pharmacologic inhibition of site 1 protease activity inhibits sterol regulatory element-binding protein processing and reduces lipogenic enzyme gene expression and lipid synthesis in cultured cells and experimental animals. J. Pharmacol. Exp. Ther. 2008, 326, 801–808. [Google Scholar] [CrossRef] [PubMed]
  109. Xiang, X.; Saha, A.K.; Wen, R.; Ruderman, N.B.; Luo, Z. AMP-activated protein kinase activators can inhibit the growth of prostate cancer cells by multiple mechanisms. Biochem. Biophys. Res. Commun. 2004, 321, 161–167. [Google Scholar] [CrossRef] [PubMed]
  110. Nguyen, T.T.; Ishida, C.T.; Shang, E.; Shu, C.; Torrini, C.; Zhang, Y.; Bianchetti, E.; Sanchez-Quintero, M.J.; Kleiner, G.; Quinzii, C.M.; et al. Activation of LXRbeta inhibits tumor respiration and is synthetically lethal with Bcl-xL inhibition. EMBO Mol. Med. 2019, 11, e10769. [Google Scholar] [CrossRef]
  111. Szász, I.; Koroknai, V.; Várvölgyi, T.; Pál, L.; Szűcs, S.; Pikó, P.; Emri, G.; Janka, E.; Szabó, I.L.; Ádány, R.; et al. Identification of Plasma Lipid Alterations Associated with Melanoma Metastasis. Int. J. Mol. Sci. 2024, 25, 4251. [Google Scholar] [CrossRef]
  112. Poli, G.; Lapillo, M.; Jha, V.; Mouawad, N.; Caligiuri, I.; Macchia, M.; Minutolo, F.; Rizzolio, F.; Tuccinardi, T.; Granchi, C. Computationally driven discovery of phenyl(piperazin-1-yl)methanone derivatives as reversible monoacylglycerol lipase (MAGL) inhibitors. J. Enzyme Inhib. Med. Chem. 2019, 34, 589–596. [Google Scholar] [CrossRef] [PubMed]
  113. Dai, S.; Yan, Y.; Xu, Z.; Zeng, S.; Qian, L.; Huo, L.; Li, X.; Sun, L.; Gong, Z. SCD1 Confers Temozolomide Resistance to Human Glioma Cells via the Akt/GSK3beta/beta-Catenin Signaling Axis. Front. Pharmacol. 2017, 8, 960. [Google Scholar] [CrossRef] [PubMed]
  114. Nashed, M.; Chisholm, J.W.; Igal, R.A. Stearoyl-CoA desaturase activity modulates the activation of epidermal growth factor receptor in human lung cancer cells. Exp. Biol. Med. 2012, 237, 1007–1017. [Google Scholar] [CrossRef] [PubMed]
  115. Pisanu, M.E.; Maugeri-Saccà, M.; Fattore, L.; Bruschini, S.; De Vitis, C.; Tabbì, E.; Bellei, B.; Migliano, E.; Kovacs, D.; Camera, E.; et al. Inhibition of Stearoyl-CoA desaturase 1 reverts BRAF and MEK inhibitioninduced selection of cancer stem cells in BRAF-mutated melanoma. J. Exp. Clin. Cancer Res. 2018, 37, 318. [Google Scholar] [CrossRef] [PubMed]
  116. Lin, H.; Patel, S.; Affleck, V.S.; Wilson, I.; Turnbull, D.M.; Joshi, A.R.; Maxwell, R.; Stoll, E.A. Fatty acid oxidation is required for the respiration and proliferation of malignant glioma cells. Neuro-Oncology 2016, 19, 43–54. [Google Scholar] [CrossRef] [PubMed]
  117. Rodríguez-Enríquez, S.; Hernández-Esquivel, L.; Marín-Hernández, A.; El Hafidi, M.; Gallardo-Pérez, J.C.; Hernández-Reséndiz, I.; Rodríguez-Zavala, J.S.; Pacheco-Velázquez, S.C.; Moreno-Sánchez, R. Mitochondrial free fatty acid b-oxidation supports oxidative phosphorylation and proliferation in cancer cells. Int. J. Biochem. Cell Biol. 2015, 65, 209–221. [Google Scholar] [CrossRef]
  118. Deng, F.; Ma, Y.X.; Liang, L.; Zhang, P.; Feng, J. The pro-apoptosis effect of sinomenine in renal carcinoma via inducing autophagy through inactivating PI3K/AKT/mTOR pathway. BioMed Pharmacother. 2018, 97, 1269–1274. [Google Scholar] [CrossRef] [PubMed]
  119. Aguirre-Portolés, C.; Feliu, J.; Reglero, G.; Ramírez de Molina, A. ABCA1 overexpression worsens colorectal cancer prognosis by facilitating tumour growth and caveolin-1-dependent invasiveness, and these effects can be ameliorated using the BET inhibitor apabetalone. Mol. Oncol. 2018, 12, 1735–1752. [Google Scholar] [CrossRef] [PubMed]
  120. Georgila, K.; Vyrla, D.; Drakos, E. Apolipoprotein A-I (ApoA-I), Immunity, Inflammation and Cancer. Cancers 2019, 11, 1097. [Google Scholar] [CrossRef]
  121. Ohmoto, T.; Nishitsuji, K.; Yoshitani, N.; Mizuguchi, M.; Yanagisawa, Y.; Saito, H.; Sakashita, N. K604, a specific acylCoA:cholesterol acyltransferase 1 inhibitor, suppresses proliferation of U251MG glioblastoma cells. Mol. Med. Rep. 2015, 12, 6037–6042. [Google Scholar] [CrossRef]
  122. LaPensee, C.R.; Mann, J.E.; Rainey, W.E.; Crudo, V.; Hunt, S.W.; Hammer, G.D. ATR-101, a Selective and Potent Inhibitor of Acyl-CoA Acyltransferase 1, Induces Apoptosis in H295R Adrenocortical Cells and in the Adrenal Cortex of Dogs. Endocrinology 2016, 157, 1775–1788. [Google Scholar] [CrossRef] [PubMed]
  123. Mehta, A.; Ratre, Y.K.; Soni, V.K.; Shukla, D.; Sonkar, S.C.; Kumar, A.; Vishvakarma, N.K. Orchestral role of lipid metabolic reprogramming in T-cell malignancy. Front. Oncol. 2023, 13, 1122789. [Google Scholar] [CrossRef] [PubMed]
  124. Deng, M.; Cao, J.; van der Kroft, G.; van Dijk, D.P.; Aberle, M.R.; Grgic, A.; Neumann, U.P.; Wiltberger, G.; Balluff, B.; Schaap, F.G.; et al. Inflammation-associated intramyocellular lipid alterations in human pancreatic cancer cachexia. J. Cachexia Sarcopenia Muscle, 2024. [Google Scholar] [CrossRef]
  125. Sieminska, J.; Miniewska, K.; Mroz, R.; Sierko, E.; Naumnik, W.; Kisluk, J.; Michalska-Falkowska, A.; Reszec, J.; Kozlowski, M.; Nowicki, L.; et al. First insight about the ability of specific glycerophospholipids to discriminate non-small cell lung cancer subtypes. Front. Mol. Biosci. 2024, 11, 1379631. [Google Scholar] [CrossRef] [PubMed]
  126. Chen, Y.; Gu, Y.; Rong, J.; Xu, L.; Huang, X.; Zhu, J.; Chen, Z.; Mao, W. Plasma-based lipidomics reveals potential diagnostic biomarkers for esophageal squamous cell carcinoma: A retrospective study. PeerJ 2024, 12, e17272. [Google Scholar] [CrossRef]
Figure 1. Summary of classes, subclasses and molecular species in glycerophospholipids.
Figure 1. Summary of classes, subclasses and molecular species in glycerophospholipids.
Metabolites 14 00312 g001
Figure 2. Experimental workflow of lipid mediators analysis using MRM HR.
Figure 2. Experimental workflow of lipid mediators analysis using MRM HR.
Metabolites 14 00312 g002
Table 1. Examples of lipids categories.
Table 1. Examples of lipids categories.
CategoriesClassesExample
Fatty acyls, FAFatty acids: straight-chain fatty acids, Eicosanoids, fatty alcohols, fatty esters, fatty amidesMetabolites 14 00312 i001Hexadecanoic acid
Prenol lipids, PRIsoprenoids, Quinones and hydroquinones, PolyprenolsMetabolites 14 00312 i0022E,6E-Farnesol
Glycerolipids, GLMonoradylglycerols: monoacyl glycerols, Diradylglycerols: diacyl glycerols, Triradylglycerols: triacyl glycerolsMetabolites 14 00312 i0031-Hexadecanoyl-2-(9Z-octadecenoyl)-sn-glycero
Glycerophospholipids, GPGlycerophosphocholines, glycerophosphoenolamines, glycerophosphoserines, glycerophospholycerols, glycerophosphoglycerophosphates, glycerophosphoinositols, glycerophosphoglycerophosphoglycerolsMetabolites 14 00312 i0041-Hexadecanoyl-2-(9Z-octadecenoyl)-sn-glycero-3-phosphocholine
Sphingolipids, SPSphingoid bases, ceramides, phosphosphingolipids, neutral glycosphingolipids, acidic glycosphingolipidMetabolites 14 00312 i005N-(Tetradecanoyl)-sphing-4-enine
Saccharolipids, SLAcylaminosugars, Acylaminosugar glycans, Acyltrehaloses, Acyltrehalose glycansMetabolites 14 00312 i006UDP-3-O-(3R-Hydroxyl-tetradecanoyl)-αD -N-acetylglucosamine
Sterol lipids, STSterols, cholesterol and derivatives, steroids, bile acids and derivativesMetabolites 14 00312 i007Cholest-5-en-3β-ol
Polyketides, PKMacrolide polyketides, Aromatic polyketides, Nonribosomal peptide/polyketides hybridsMetabolites 14 00312 i008Aflatoxin B1
Adapted with modifications from [11].
Table 2. Major functions of lipid classes.
Table 2. Major functions of lipid classes.
Membrane componentsGlycerophospholipids (e.g., PC, PE, PI, PS, PG, PA, etc.), sphingolipids (e.g., sphingomyelin, cardiolipin, cerebroside, sulfatide, gangliosides, etc.), glycolipids, sterol lipids (e.g., cholesterol, etc.)
Energy storage and metabolismGlycerolipids (e.g., NEFA, TAG, diacylglycerol, monoacylglycerol, acyl CoA, acylcarnitine, etc.)
SignalingGlycerolipids (e.g., diacylglycerol, monoacylglycerol, acyl CoA, acylcarnitine, NEFA, oxidized fatty acid), sphingolipids (e.g., Cer, sphingosine, sphingosine-1-phosphate), psychosine, N-acylethanolamine, Lyso-lipids, etc.
Other functionsPlasmalogen (antioxidant)
Acylcarnitine (lipid transport)
CL (mitochondrial respiration),
PS (cofactor, substrate of PE synthesis)
Choline glycerophospholipid (PC), ethanolamine glycerophospholipid (PE), phosphatidylinositol (PI), phosphatidylglycerol (PG), phosphatidic acid (PA), phosphatidylserine (PS), sphingomyelin (SM), cardiolipin (CL), Galactosylceramide (GalCer), Glucosylceramide (GluCer), non-esterified fatty acid (NEFA), triacylglycerol (TAG), diacylglycerol (DAG), monoacylglycerols (MAG), Eicosanoids and other oxidized fatty acids, sphingoid-1-phosphate (S1P).
Table 4. Lipidomic studies on biofluids reporting diagnostic lipid signatures.
Table 4. Lipidomic studies on biofluids reporting diagnostic lipid signatures.
CancerLipids Modified Technique
OvarianPPE(16:0, 18:1), LPC15:0 LPA18:2, PPE(18:0, 22:6)LC/ESI/MS/MS
LiverLPC16:0 LPC18:0 PS16:0, PC18:0UPLC-MS
Liver18:2n-6, 18:1n-9, 20:4n, 16:0GC-MS
LungSM(16:0, 16:1), LPC18:1 LPC20:4 LPC20:3 LPC22:6FTICR-MS
LungLPE18:1 CE18:2 ePE40:4, SM22:0ESI-MS
ProstateePC38:5 PC40:3, PC42:4Q-TOF-MS
BreastPalmitic acid, 16:0, Stearic acid 18:0, Linoleic acid, 18:2, Alpha-Linolenic acid, 18:3, Eicosapentaenoic acid 20:5, Docosapentaenoic acid 22:5GC-MS
ColorectalCapric acid, 10:0GC-MS
Adapted with modifications from [46].
Table 5. Potential lipid biomarkers in the oncology field.
Table 5. Potential lipid biomarkers in the oncology field.
TumorLipid ClassLipid TypeSampleUpDownRef.
Lung cancer
(adenocarcinomas)
GPLsPI arachidonate containing phospholipidsTissueX [47]
Lung cancer
(adenocarcinomas)
GPLsPC 32:0, PC 32:1, PGsTissue X[47]
Lung cancer
(adenocarcinomas)
FAsfree arachidonicTissueX [47]
Lung cancer
(NSCLC)
GPLsPC 34:1, PC 36:2, PC 36:3TissueX [48]
Lung cancer
(NSCLC)
GPLs, SLsPC 32:0, ST-OH 42:1TissueX [48]
Lung cancer
(NSCLC)
FAsEPATissueXX[49]
Lung cancer
(Different cancer type)
GPLsSM 16:0/1, LPC 18:1, LPC 20:4, LPC 20:3, LPC 22:6SerumX [50]
Lung cancer
(NSCLC)
GPLsPI 38:3, PI 40:3, PI 38:2TissueX [51]
Lung cancer
(NSCLC)
SLsSM 40:1, SM 42:1, SM 36:1Tissue X[51]
Breast cancer NSGPLs, SLsPC, PE, PI, SMsTissueX [52]
Breast cancer
(luminal, HER2+ and triple-negative)
GPLsPC 34:1TissueX [53]
Breast cancer (MDA-MB 231 model)Palmitoyl carnitines, stearoyl carnitine GPLs, SLsPC16:0/16:0,PC16:0/18:1, PC18:1/18:1,PC18:0/18:1 PC16:0/22:1 SMd18:1/16:0TissueX [54]
Breast cancer NSGPLsPI18:0/20:4Urine X[55]
Breast cancer NSGPLsPS (18:1/18:1 18:2/18:0)UrineX [55]
Breast cancer NSGPLsPI 18:0/18:1, PI 18:0/20:3TissueX [56]
Breast cancer
(mammary epithelial and breast cancer)
GPLsPCs, PI 22:5/18:0, PI 18:0/18:1Cell lineXX[57]
Breast cancer NSGPLsPS 18:0/20:4, PI 18:0/20:4, PC 18:0/20:4Cell lineX [58]
Breast cancer NSGPLsPIs, Pes, PCs, LPCsTissueXX[59]
Breast cancer NSGPLsTGs containing C18:1 fatty acyl chainSerum X[60]
Breast cancer NSFAsLinoleic acid (C18:2)Serum X[61]
Prostate cancer NSGPLsPS 18:0/18:1, PS 16:0/22:6UrineX [62]
Prostate cancer NSGPLsPS 18:1/18:0, PS 18:0/20:5Urine X[62]
Prostate cancer NSGPLsPI 18:0/18:1, PI 18:0/20:3, PI 18:0/20:2TissueX [63]
Localized prostate cancerGPLsLPC 16:0/OH, SM d18:1/16:0Tissue X[64]
Localized prostate cancerGPLsLPC 16:0/OHTissue X[64]
Localized prostate cancerGPLsPC 40:3, PC 42:4SerumX [65]
Prostate cancer NSGPLsPC 39:6SerumX [66]
Prostate cancer NSFAsFA 22:3SerumX [66]
Colorectal cancer NSGPLsLPC 18:1, LPC 18:2Plasma X[67]
Colorectal cancer (pT3 stage, various grades (G2, G3))GPLsPC/PE ratioCell linesX [68]
Colorectal cancer NSGPLsPC 16:0/16:1TissueX [69]
Colorectal cancer NSGPLsPC 16:0/18:1, LPC 16:0, LPC 18:1TissueX [70]
Colorectal cancer liver metastasisGPLsPE 38:6, PE 40:4TissueX [71]
Colorectal cancer NSFAsn-3 PUFAsRed Blood cell X[72]
Colorectal cancer NSFAsn-6-PUFA/n-3-PUFARed Blood cellX [72]
Ovarian cancer NSGPLsLPCPlasmaX [73]
Ovarian cancer NSGPLsPC, TGPlasma X[73]
Epithelial ovarian cancerGPLsTGs 50:2 50:152:2 54:4 54:3Cell linesX [74]
Ovarian cancer NSGPLsPC 32:3, PC 34:1, PC 36:2TissueX [75]
Ovarian cancer NSGPLsLPA 16:0, LPA 20:4PlasmaX [76]
Ovarian cancer and other gynecological cancersGPLsLPASerum/plasmaX [77]
Ovarian cancer and other gynecological cancersGPLsLPA 16:0, LPA 18:2, LPA18:1, LPA18:0, LPI 16:0, LPI 18:0, LPI 20:4PlasmaX [78]
Benign and malignant ovarian cancerGPLsLPAPlasmaX [79]
Benign and malignant ovarian cancerGPLsLPAPlasmaX [80]
Benign and malignant ovarian cancerGPLsPlasmalogen phospatidylethanol, PC, plasmalogen PC, SM and LPC X [81]
Metastatic pancreatic cancerSLsCeramides species (C16:0 and C24:1)Tissue/PlasmaX [82]
Metastatic pancreatic cancerSLsC18:0 C20:0 C22:0 C24:0 C24:1Tissue/Plasma X[82]
Metastatic pancreatic cancerSLsC16:0 C20:0 C22:0 C24:0 C24:1Tissue/PlasmaX [82]
Pancreatic cancer PANC-1 cellsGPLsLPACell linesX [83]
Pancreatic cancer NSFAsMUFAPlasmaX [84]
Gastric cancer NSGPLsPC16:0/18:1TissueX [85]
Gastric cancer NSGPLsLPC 16:0Tissue X[85]
Bladder Cancer NSGPLsPS 18:0/18:1TissueX [86]
Bladder CancerGPLsPI 18:0/20:4TissueX [86]
Bladder Cancer (Model of human invasive bladder cancer)GPLsPS 18:0/18:1TissueX [87]
Bladder Cancer (Model of human invasive bladder cancer)GPLsPG 18:1/18:1TissueX [87]
Bladder Cancer (Model of human invasive bladder cancer)GPLsPI 16:0/18:1TissueX [87]
Bladder Cancer (Model of human invasive bladder cancer)GPLsPI 18:0/18:1TissueX [87]
Bladder Cancer (Model of human invasive bladder cancer)GPLsPS 18:1/18:1TissueX [87]
Esophageal cancer (ESCC)GPLsOctanoylcarnitine, LPC 16:1, DecanoylcarnitinePlasmaX [88]
Esophageal cancer (OSCC)GPLsPC 16:0/16:1TissueX [89]
Esophageal cancer (OSCC)GPLsPC 18:1/20:4Tissue X[89]
Esophageal cancer (ESCC)GPLsPS, PA, PC, PI, PEPlasmaX [90]
Kidney cancer NSGLs SLsPE (P-16:0e/0:0), ganglioside GM3, (d18:1/22:1), sphinganine C17, SMd18:0/16:1(9Z)SerumX [91]
Kidney cancer NSGPLs STLs
Gls
PC, Plasmalogens, cholesterol esters, TGsTissueX [92]
Kidney cancer NSGPLs, FAsPE, Unsaturated FAsTissue X[92]
Kidney cancer NSGPLsPL, PE 36:1, PC 38:4, PC 36:2, PC 32:0TissueX [93]
Kidney cancer NSGPLsPE 34:2, PE 36:4, PE 38:4, PC 34:1 PC34:2 PC 36:4 PI36:4Tissue X[93]
Kidney cancer NSGPLsPI 18:0/20:4, PI22:4/18:0 PS18:0/18:1 PG18:1/18:1TissueX [94]
Kidney cancer (Human papillary renal carcinoma)FAsFA12:0 Tissue X[94]
Thyroid cancer (Thyroid papillary cancer)GPLsPC 16:0/18:1 PC 16:0/18:2TissueX [95]
Thyroid cancer (Thyroid papillary cancer)SLsSMd18:0/16:1TissueX [95]
Malignant and benignant thyroid cancerGPLsPC 34:1, PC 36:1, PC 32:0Tissue/SerumX [96]
Malignant and benignant thyroid cancerGPLsPA 36:62, PA 36:3, PA 38:4, PA 38:5, PA 40:5Tissue/SerumX [96]
Adapted with modifications from reference [97].
Table 6. Overview of potential agents as modulators of lipid metabolism in cancer.
Table 6. Overview of potential agents as modulators of lipid metabolism in cancer.
TargetAgent NameType of Cancer/Phase of DevelopmentComments and Benefits Reference/Clinical Trial
Fatty acid synthase (FASN) inhibitorOrlistatProstate, breast, ovarian, colon cancer and other solid tumors/ApprovedAntiobesity lipase and FASN inhibitor; not yet assessed clinically for cancer. An anti-obesity drug approved by the FDA and an irreversible inhibitor of FASN[98]
TVB-3166/TVB-3664Oral squamous cell carcinoma, colorectal, breast cancer/PreclinicalA reversible and selective FASN inhibitor[99,100]
Conjugated Linoleic AcidBreast Cancer/Phase IReduce FASN gene expression and spot 14NCT00908791
OmeprazoleTriple-negative breast cancer/Phase IIA proton pump inhibitors that can inhibit FASNNCT02595372
IPI-9119Castration-resistant prostate cancer/PreclinicalSelectively inhibit FASN and suppress expression of both full-length of androgen receptor (AR) and AR variant V7[101]
TVB-2640Solid Malignant Tumor/Phase IITrials in breast cancer +/− trastuzumab/docetaxel. A potent and reversible FASN inhibitorNCT03032484
NCT03179904
Acetyl-CoA Carboxylase (ACC)TOFALung cancer and colon carcinoma/PreclinicalInduce apoptosis as an allosteric inhibitor of ACC-alpha[102]
Soraphen AProstate cancer/PreclinicalInhibit fatty acid synthesis and stimulate fatty acid oxidation[103]
ND-646NSCLC/PreclinicalInhibit fatty acid synthesis and tumor growth as an allosteric inhibitor of the ACC[104]
3-hydroxy-3-methylglutaryl coenzyme A reductase (HMGCR)Statins, e.g., lovastatin,
atorvastatin, rosuvastatin
and simvastatin
Various leukemia and solid tumors/Phase IIInhibitors of HMGCR. Association studies between statin use and outcomes in many cancers, some randomized trialsNCT03358017 NCT03324425 NCT02569645 NCT03275376
CD36ABT-510Prostate cancer/Phase IITrial in metastatic melanoma. Reduced fatty acid uptake and the abundance of oncogenic signaling lipids[105]
SREBPsFatostatinProstate cancer/PreclinicalA non-sterol diarylthiazole derivative which has antimitotic properties and perturbs nuclear translocation of SREBP and androgen receptor signaling[106]
BetulinBRAFV600E-mutant melanoma/PrelinicalIncrease membrane lipid poly-unsaturation and lipid peroxidation; sensitize therapy-resistant melanoma cells to MAPK-targeting therapy[107]
PF-429242HCC/PreclinicalReversible site 1 protease inhibitors, which inhibit endogenous SREBP processing[108]
AMPKMT 63–78Prostate cancer/PreclinicalMT 63–78 is a specific and potent direct AMPK activator and also induces cell mitotic arrest and apoptosis[109]
5-aminoimidazole 4-
carboxamide riboside (AICAR)
Prostate cancer/PreclinicalPreclinical studies in cancer and an AMPK activator which inhibits cell growth[109]
LXRLXR-623Phase I for atherosclerosisPreclinical studies in cancer[110]
GW3965; LXR623Melanoma and glioblastoma/PreclinicalLXR agonists which suppress mitochondrial respiration and decrease cholesterol levels by enhancing the excretion and decreasing the resorption of cholesterol[111]
ATGLatglistatin Specific for murine ATGL
MAGLJZL184Cancer-associated bone disease/PreclinicalPotent and reversible MAGL inhibitors[112]
SCDA939572Glioblastoma and renal cell carcinoma/PreclinicalInhibit tumor growth both in vitro and in vivo; overcoming chemotherapy agent resistance[113]
CVT-11127 or CVT- 12012 Lung cancer/PreclinicalA small molecule SCD inhibitor which modulates cancer cell metabolism, proliferation and survival[114]
MF-438Lung cancer/PreclinicalInduce lung cancer stem cell apoptosis[115]
CPT1EtomoxirGlioma/PreclinicalA CPT1 inhibitor which inhibits proliferative activity[116]
PerhexilineBreast and gastrointestinal cancer/PreclinicalA CPT1 inhibitor which blocks FFA utilization, OxPhos and proliferation[117]
PLDST1326Preclinical
FIPIPreclinicalTargets immune infiltration into tumors
VU0155072–2PreclinicalTargets immune infiltration into tumors
FABPEI-05Preclinical
SBFI-102, 103Preclinical
Choline kinaseHC-3, JCR and MN
derivatives
Preclinical
Triterpene quinone methidesPreclinicalDerived from natural products
ICL-CCIC-0019Preclinical
EB-3D, EB-3PPreclinical
FarnesyltransferasetipifarnibPhase I/IIPhase II studies in metastatic BC and squamous cell carcinoma; Phase I trial in glioblastoma
lonafarnibPhase I/IbLargely combination studies in glioblastoma, BC
BMS-214662Phase ISingle agent or in combination with paclitaxel in advanced solid tumors
Palmitoylation2-bromopalmitatePreclinical
Acid sphingomyelinaseFluphenazinePreclinical
Mammalian target of rapamycin (mTOR)RapamycinBreast cancer/PreclinicalInhibited S6 phosphorylation and cell proliferation, and resulted in lower levels of apoptosis induction[118]
EverolimusCastrated Resistant Prostate Cancer. Locally Advanced Cervical Cancer/Phase IIIDirectly inhibit mTORC1 and indirectly inhibit mTORC2NCT03580239
PF-05212384Advanced Cancer; Advanced squamous cell lung, pancreatic, head and neck, and other solid tumors/Phase IIntravenous PI3K/mTOR inhibitorNCT01347866
NCT03065062
PF-04691502Breast Neoplasms/Phase IIInhibit PI3K and mTOR kinaseNCT01658176
Vistusertib/AZD2014Endometrial, triple negative breast cancer, ovarian, primary peritoneal, or fallopian tube cancer/Phase IImTORC1/2 InhibitorNCT02208375
Protein kinase B (PKB), AktMK-2206Advanced or metastatic solid tumors or breast cancer; prostate cancer/Phase I, IIInhibit Akt phosphorylation, cell proliferation and apoptosis in a dose-dependent manner NCT01245205
NCT01277757
NCT01251861
Capivasertib/AZD5363Breast cancer, prostate cancer and advanced solid tumors/Phase I, IIA novel pan-AKT kinase catalytic inhibitorNCT03310541
NCT02525068
GSK2141795Endometrial cancer/Phase IAKT inhibitorNCT01935973
ApoA-IApabetalone (RVX-208)Colorectal cancer/PreclinicalA BET inhibitor which is a stimulator of ApoA-I and regulates the reverse cholesterol transport[119]
ApoA-1 mimetic peptidesPreclinicalMimetic peptides which are synthesized on the basis of aamphipathic helical repeating structure of ApoA-I[120]
ACATK604Glioblastoma/PreclinicalA selective ACAT1 inhibitor, which suppresses proliferation of glioblastoma cells[121]
ATR-101Advanced adrenocortical carcinoma/Phase IAn oral selective ACAT inhibitorNCT01898715
[122]
Table 7. Lipid metabolism in functional subsets of T cells.
Table 7. Lipid metabolism in functional subsets of T cells.
T-Cell SubsetMetabolic EventsMolecular Players
Naïve T cellMajorly rely on OXPHOS and FAOEnzymes of OXPHOS and FAO
Ag- stimulated
T cells
Aerobic glycolysis, FA synthesis, uptake, and accumulationmTOR, PI3K, c-myc, HIF, GLUTs, FASN, SREBP, and PPARg
CD8+ cytotoxic T cellsMainly rely on FA and lipid synthesisFASN, ACC1, HMGCR, and SREBP
CD4+ helper T cells FA and lipid synthesis and FA uptakeASN, SREBPs, CD36, FABPs, GPR43, GLP84, and LXR
Th1Relatively high FA uptake dominates over FA synthesisCD36, FABPs, FASN, and LCFA favor Th1 activation
Th2Relatively low FA uptake, however, dominates over FA synthesisCD36, FABPs, FASN, and PUFA favor Th2 activation
Th3Profound FA synthesis and aerobic glycolysisFASN, ACC1, PDHK, LXR, and 2-HG favor Th17 activation
Regulatory T cells
(Treg)
Exogenous FA uptake dominates over FA synthesis, OXPHOS, and
FAO
Foxp3, FABP5, CPT1, downregulation of ACC1, SREBPs, and SCAP
Memory T cells
(Tm)
OXPHOS and FAO, FA uptake, and downregulation of FA synthesisEnzymes of OXPHOS and FAO, FABP4/5, AMPK, and downregulation of
ACC1
OXPHOS, oxidative phosphorylation; FAO, fatty acid oxidation; FA, fatty acid. Adopted from reference [122].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nassar, A.F.; Nie, X.; Zhang, T.; Yeung, J.; Norris, P.; He, J.; Ogura, H.; Babar, M.U.; Muldoon, A.; Libreros, S.; et al. Is Lipid Metabolism of Value in Cancer Research and Treatment? Part I- Lipid Metabolism in Cancer. Metabolites 2024, 14, 312. https://doi.org/10.3390/metabo14060312

AMA Style

Nassar AF, Nie X, Zhang T, Yeung J, Norris P, He J, Ogura H, Babar MU, Muldoon A, Libreros S, et al. Is Lipid Metabolism of Value in Cancer Research and Treatment? Part I- Lipid Metabolism in Cancer. Metabolites. 2024; 14(6):312. https://doi.org/10.3390/metabo14060312

Chicago/Turabian Style

Nassar, Ala F., Xinxin Nie, Tianxiang Zhang, Jacky Yeung, Paul Norris, Jianwei He, Hideki Ogura, Muhammad Usman Babar, Anne Muldoon, Stephania Libreros, and et al. 2024. "Is Lipid Metabolism of Value in Cancer Research and Treatment? Part I- Lipid Metabolism in Cancer" Metabolites 14, no. 6: 312. https://doi.org/10.3390/metabo14060312

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop