Next Article in Journal
Autophagy Regulators in Cancer
Next Article in Special Issue
Humanization of the Reaction Specificity of Mouse Alox15b Inversely Modified the Susceptibility of Corresponding Knock-In Mice in Two Different Animal Inflammation Models
Previous Article in Journal
Systemic Inflammation after Aneurysmal Subarachnoid Hemorrhage
Previous Article in Special Issue
The Sizes and Composition of HDL-Cholesterol Are Significantly Associated with Inflammation in Rheumatoid Arthritis Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Inhibitory Investigations of Acyl-CoA Derivatives against Human Lipoxygenase Isozymes

1
Department of Chemistry and Biochemistry, University of California Santa Cruz, Santa Cruz, CA 95064, USA
2
Department of Biochemistry and Pharmacology, University of Melbourne, Melbourne, VIC 3010, Australia
3
Walter and Eliza Hall Institute of Medical Research, Parkville, VIC 3052, Australia
4
Department of Medical Biology, University of Melbourne, Melbourne, VIC 3010, Australia
5
Drug Discovery Biology, Monash Institute of Pharmaceutical Sciences, Monash University, Parkville, VIC 3052, Australia
6
ARC Centre for Cryo-Electron Microscopy of Membrane Proteins, Monash Institute of Pharmaceutical Sciences, Monash University, Parkville, VIC 3052, Australia
7
Department of Pharmacology, University of Michigan Medical School, Ann Arbor, MI 48109, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 10941; https://doi.org/10.3390/ijms241310941
Submission received: 5 June 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 30 June 2023
(This article belongs to the Special Issue Lipid Signaling and Metabolism in Inflammation-Associated Diseases)

Abstract

:
Lipid metabolism is a complex process crucial for energy production resulting in high levels of acyl-coenzyme A (acyl-CoA) molecules in the cell. Acyl-CoAs have also been implicated in inflammation, which could be possibly linked to lipoxygenase (LOX) biochemistry by the observation that an acyl-CoA was bound to human platelet 12-lipoxygenase via cryo-EM. Given that LOX isozymes play a pivotal role in inflammation, a more thorough investigation of the inhibitory effects of acyl-CoAs on lipoxygenase isozymes was judged to be warranted. Subsequently, it was determined that C18 acyl-CoA derivatives were the most potent against h12-LOX, human reticulocyte 15-LOX-1 (h15-LOX-1), and human endothelial 15-LOX-2 (h15-LOX-2), while C16 acyl-CoAs were more potent against human 5-LOX. Specifically, oleoyl-CoA (18:1) was most potent against h12-LOX (IC50 = 32 μM) and h15-LOX-2 (IC50 = 0.62 μM), stearoyl-CoA against h15-LOX-1 (IC50 = 4.2 μM), and palmitoleoyl-CoA against h5-LOX (IC50 = 2.0 μM). The inhibition of h15-LOX-2 by oleoyl-CoA was further determined to be allosteric inhibition with a Ki of 82 +/− 70 nM, an α of 3.2 +/− 1, a β of 0.30 +/− 0.07, and a β/α = 0.09. Interestingly, linoleoyl-CoA (18:2) was a weak inhibitor against h5-LOX, h12-LOX, and h15-LOX-1 but a rapid substrate for h15-LOX-1, with comparable kinetic rates to free linoleic acid (kcat = 7.5 +/− 0.4 s−1, kcat/KM = 0.62 +/− 0.1 µM−1s−1). Additionally, it was determined that methylated fatty acids were not substrates but rather weak inhibitors. These findings imply a greater role for acyl-CoAs in the regulation of LOX activity in the cell, either through inhibition of novel oxylipin species or as a novel source of oxylipin-CoAs.

1. Introduction

Fat and lipid metabolism is a complex process that involves the synthesis, breakdown, and transport of lipids throughout the body [1]. Lipids not only play a vital role in providing energy for the body but also help in the formation of cell membranes [2], hormone synthesis [3,4], and storage of fat-soluble vitamins [5]. Dysregulation of lipid metabolism has been associated with a range of diseases, including obesity [6,7], type 2 diabetes [8,9], and cardiovascular disease [10,11]. One important aspect of lipid metabolism is the synthesis and breakdown of fatty acids, the building blocks of triglycerides, phospholipids, and cholesterol esters [12,13]. Fatty acid metabolism involves a series of enzymatic reactions that convert fatty acids into acetyl-CoA, which can then enter the citric acid cycle for energy production [14,15,16]. In addition to their role in energy production, fatty acid-CoA esters have been shown to act as signaling molecules [17,18], regulating the activity of enzymes involved in lipid metabolism and other cellular processes [19,20]. Some studies have suggested that acyl-CoA synthase can even modulate the activity of cyclooxygenase (COX) and lipoxygenase (LOX) isozymes, which are involved in the production of pro-inflammatory lipid mediators [21]. For example, it has been observed that COX can be inhibited by palmitoyl-CoA, which can have concentrations as high as 10 µM in platelets [22]. Since COX reacts with fatty acids, such as arachidonic acid (AA), to produce specialized pro-resolving lipid mediators (SPMs) [23,24], this mechanism of inhibition could have significant biological effects. Lipoxygenase is also involved in the production of SPMs and uses the same substrate profile to produce potent pro- and anti-inflammatory mediators involved in allergic reactions and other immune responses [25,26]; therefore, inhibition of LOX isozyme by acyl-CoAs could have significant biological effects.
Lipoxygenases are a class of non-heme iron-containing dioxygenases that catalyze the oxygenation of polyunsaturated fatty acids, resulting in the formation of lipid mediators such as leukotrienes and lipoxins [27,28,29]. The high-resolution cryo-EM structures of human 12S-lipoxygenase (h12-LOX) were determined for the first time in a recent study, providing valuable insights into its oligomeric states, but equally interesting was the discovery that an acyl-CoA was bound to its active site [30]. The resolution of the h12-LOX structure was not high enough to identify the acyl-CoA; however, screening a family of acyl-CoAs showed that oleoyl-CoA was the most likely candidate due to its potency (IC50 = 32 µM). The current work expands on this result, screening a larger family of acyl-CoAs against h5-LOX, h12-LOX, h15-LOX-1, and h15-LOX-2. This work provides new insights into the inhibitory mechanism of acyl-CoAs against LOX isozymes, potentially indicating an endogenous role for acyl-CoAs as LOX inhibitors, which could lead to the development of novel therapeutics targeting these enzymes.

2. Results and Discussion

2.1. IC50 Determination of Unsaturated Fatty Acyl-CoAs

For h12-LOX, the IC50 values for palmitoyl-CoA (16:0), palmitoleoyl-CoA (16:1), stearoyl-CoA (18:0), linoleoyl-CoA (18:2), ɣ-linolenoyl-CoA (18:3), and docosahexaenoyl-CoA (22:6) indicated weak inhibition, with values greater than 200 µM (Table 1). Oleoyl-CoA (18:1) and arachidonoyl-CoA (20:4) had slightly greater inhibition potency, with IC50 values of 32 ± 4 µM and 110 ± 20 µM, respectively (Table 1). These data have been published previously [30].
For h15-LOX-1, the IC50 values for palmitoyl-CoA (16:0), palmitoleoyl-CoA (16:1), linoleoyl-CoA (18:2), ɣ-linolenoyl-CoA (18:3), arachidonoyl-CoA (20:4), and docosahexaenoyl-CoA (22:6) indicated weak inhibition, with all values being greater than 50 µM (Table 1). However, oleoyl-CoA (18:1) and stearoyl-CoA (18:0) had relatively stronger inhibition, with values of 39 ± 2 µM and 4.2 ± 0.6 µM, respectively. The data indicate that C18 acyl-CoAs have the greatest potency; however, since ɣ-linolenoyl-CoA (18:3) is not inhibitory, the degree of saturation also affects potency.
For h15-LOX-2, the IC50 values for palmitoyl-CoA (16:0), palmitoleoyl-CoA (16:1), linoleoyl-CoA (18:2), ɣ-linolenoyl-CoA (18:3), arachidonoyl-CoA (20:4), and docosahexaenoyl-CoA (22:6) indicated weak inhibition, with values greater than 100 μM (Table 1). Stearoyl-CoA (18:0) had relatively strong inhibition, with a value of 7.6 ± 1 µM; however, oleoyl-CoA (18:1) had exceptionally strong inhibition, with an IC50 of 620 ± 60 nM. Similar to what was seen for h15-LOX-1, the C18 acyl-CoAs had the greatest potency; however, with the presence of three double bonds, as with ɣ-linolenoyl-CoA (18:3), all inhibitor potency was lost.
For h5-LOX, the IC50 values for ɣ-linolenoyl-CoA (18:3) and docosahexaenoyl-CoA (22:6) indicated weak inhibition, with values greater than 200 µM. Stearoyl-CoA (18:0), oleoyl-CoA (18:1), and linoleoyl-CoA (18:2) had slightly stronger inhibition, with IC50 values greater than 50 µM (Table 1). However, palmitoyl-CoA (16:0) and palmitoleoyl-CoA (16:1) had low micromolar inhibition, with values of 3.3 ± 0.3 µM and 2.0 ± 0.4 µM. The C16 acyl-CoAs’ potency against h5-LOX contrasts with the other LOX isozymes and is possibly due to the active site requirement for the reverse binding of the substrate of h5-LOX relative to the other isozymes.
In summary, the majority of the acyl-CoAs were not inhibitors against LOX isozymes; however, certain acyl-CoAs displayed micromolar potency. For h5-LOX, the C16 acyl-CoAs displayed the most promising inhibitory properties, with low micromolar potencies. However, with all the other isozymes—h15-LOX-1, h15-LOX-2, and h12-LOX—the C18 acyl-CoAs had the most potent inhibitory properties. Interestingly, adding double bonds to the acyl chain, such as from oleoyl-coA (18:1) to linoleoyl-CoA (18:2), eliminated significant inhibition potency, and complete loss of inhibitory activity was observed with γ-linolenoyl-CoA (18:3). This effect was most likely due to the added restriction on structural freedom with the additional double bonds, which would decrease flexibility and lead to lower binding affinity, thus reducing the availability of the proper orientation. With respect to biological relevancy, the most potent acyl-CoA interaction was that between oleoyl-CoA and h15-LOX-2, which displayed submicromolar potency. This level of potency drew us to undertake a more detailed kinetic analysis (vide infra).

2.2. Acyl-CoA Substrate Activity

Considering that some of these acyl-CoAs could potentially be substrates for LOX isozymes, each acyl-CoA was tested at 70 µM against each LOX isozyme to establish the possibility of catalysis. These measurements revealed that none of the acyl-CoAs were substrates against the LOX isozymes, except for linoleoyl-CoA (18:2) and h15-LOX-1, as monitored by a change in absorbance at 234 nm (Table 2). The kinetic parameters were determined to be 7.5 ± 0.4 s−1 for kcat, 12 ± 0.8 µM for KM, and 0.62 ± 0.1 µM−1 s−1 for kcat/KM at 22 °C. These kinetic parameters are comparable to that of free linoleic acid, with kcat being 1.3-fold less and kcat/KM being 7-fold less than the published values [31], suggesting it as a possible biological substrate for h15-LOX-1.

2.3. IC50 Determination and Substrate Activity of Methyl Ester Unsaturated Fatty Acids

The acyl-CoAs are large ester derivatives of fatty acids and, given this fact, it was deemed appropriate to determine if simpler esters could also inhibit the LOX isozymes. Thus, methyl-13Z,16Z-docosadienoic, methyl 13(Z),16(Z),19(Z)-docosatrienoate, and methyl docosatetraenoate were investigated and determined to be weak inhibitors, with IC50 values greater than 200 µM for h12-LOX, h-15LOX-1, and h15-LOX-2. It should be noted that none of the methyl ester unsaturated fatty acids were substrates, except for methyl ester arachidonic acid and h15-LOX-1, as previously observed. However, the rate of this activity was low enough to still allow for IC50 measurements.

2.4. Kinetic Investigation of Oleoyl-CoA (18:1) Inhibition of h15-LOX-2

Oleoyl-CoA displayed nanomolar IC50 potency against h15-LOX-2 (620 ± 60 nM), which led to further kinetic investigations at varying inhibitor concentrations. The replot of the apparent KM versus concentration for oleoyl-CoA exhibited a hyperbolic response with increasing amounts of acyl-CoA from 624 nM to 3.6 μM (Figure 1). This hyperbolic response is representative of allosteric inhibition (i.e., partial inhibition), which was previously observed with 13-(S)-HOTrE(γ) and h15-LOX-2, as described in Scheme 1 (Equations (1)–(4)) [32].
1 / v = ( α K M / k cat ) [ ( [ I ] + K i ) / ( β [ I ] + α K i ) ] × 1 / [ S ] + 1 / k cat × [ ( [ I ] + α K i ) / ( β [ I ] + α K i ) ]
K M ( app ) = α K M [ ( [ I ] + K i ) / ( [ I ] + K i ) ]
k cat / K M = ( k cat / α K M ) [ ( β [ I ] + α K i ) / ( [ I ] + K i ) ]
k cat = k cat [ ( β [ I ] + α K i ) / ( [ I ] + α K i ) ]
Utilizing the equations derived from Scheme 1, the KM (app) values were fitted with Equation 2 to yield a value for α of 3.2 +/− 1 and a value for Ki of 82 +/− 70 nM (Figure 1). These two parameters were then utilized to fit Equation (3) relative to kcat/KM (app), which yielded a β value of 0.26 +/− 0.23 (Figure 2). Given the high error for the β value, β was also determined by applying the values for α and Ki from the KM (app) plot to Equation 4 and fitted to the kcat data (Figure 3). This fit yielded a β value of 0.30 +/− 0.07, consistent with the Equation 3 fit but with a lower error value. These kinetic values indicate hyperbolic allostery, α > 1 (K-type inhibition) and β < 1 (V-type inhibition), with the kinetic change being equally observed in both KM (α = 3.2 +/− 1) and kcat (β = 0.30 +/− 0.07). This combined effect is emphasized in the β/α value, the allosteric effect on kcat/KM, which is significantly less than 1 (β/α = 0.09) and indicates kcat/KM allosteric inhibition. These allosteric kinetic data indicate the formation of a catalytically active ternary complex (I·E·S) between h15-LOX-2 and oleoyl-CoA and are consistent with our previous finding of an allosteric site for 13-(S)-HODE with 15-LOX-2 [32]. It should be noted that the Ki values from our hyperbolic fits (82 +/− 70 nM) were dramatically lower than those determined from IC50 measurements (620 ± 60 nM), which is consistent with the more accurate nature of the hyperbolic fits relative to the IC50 fits, due to the oversimplification of the IC50 analysis.
Interestingly, in our recently published structure [30], oleoyl-CoA was bound in the active site of 12-LOX with its head group occupying the entrance to the binding site and the fatty acid tail extending deep into the U-shaped catalytic cavity. Similar to the previously predicted pose of AA bound to h12-LOX [33], H596 forms hydrogen bonds with the carbonyl group of the thioester bond. This is also consistent with the predicted pose of AA bound to h12-LOX, where L407, F414, A417, and V418 form Van der Waals interactions with the oleic acid tail. The h12-LOX binding site is quite narrow, consisting of two mostly linear segments connected by a single kink near the C9 position of the oleic acid (equivalent to C11 of AA). The shape of the binding cavity likely restricts the binding of the other acyl-CoAs tested in this study due to the strict requirements imposed on the position and saturation state of the fatty acid. With respect to the allosteric inhibition observed for h15-LOX-2, its binding site and that of h5-LOX and h15-LOX-1 appear to be wider and shorter, providing a possible explanation for their slightly better tolerance of different fatty acid saturation states.

3. Materials and Methods

3.1. Chemicals

Fatty acids used in this study were purchased from Nu Chek Prep, Inc. (Elysian, MN, USA). All other solvents and chemicals were reagent-grade or better and were used as purchased without further purification.

3.2. Synthesis of DHA-CoA

Reactions took place in an oxygen-deficient environment. DHA was added to an equimolar amount of butylated hydroxytoluene. Oxalyl chloride, along with 1 drop of dry dimethylformamide (DMF), was then added, and the mixture was stirred for 1 h at 37 °C and dried under vacuum. CoA was then added to a solution of tetrahydrofuran and 150 mM sodium bicarbonate in a ratio of 2.2/1.0. This solution was combined with the DHA solution for 30 min at 37 °C. The solution was dried to remove the tetrahydrofuran. Dichloromethane with 1.3% perchloric acid was used to extract the desired product away from the unreacted DHA acid. The water layer was retained and mixed with an equal volume of acetonitrile and lyophilized. The solid was washed with dry acetone and ether (Scheme 2). The synthesis was based on published protocols [34,35].

3.3. Expression and Purification of h5-LOX, h12-LOX, h15-LOX-1, and h15-LOX-2

Overexpression and purification of his-tagged wild-type h12-LOX (Uniprot entry P18054) [36], h15-LOX-1 (Uniprot entry P16050) [36], and h15-LOX-2 (Uniprot entry O15296) [37] were performed using nickel-affinity chromatography. The purity of h12-LOX, h15-LOX-1, and h15-LOX-2 was assessed using SDS gel as greater than 85%, and metal content was assessed with a Finnigan inductively coupled plasma mass spectrometer (ICP-MS, Thermo Fischer Scientific, San Jose, USA) via comparison with iron standard solution. Cobalt-EDTA was used as an internal standard. WT h5-LOX (UniProt entry P09917) was expressed in Rosetta 2 cells (Novagen) transformed with the pET14b-Stable-5-LOX plasmid (a gift from Marcia Newcomer of Louisiana State University) and grown in Terrific Broth containing 34 μg/mL chloramphenicol and 100 μg/mL ampicillin at 37 °C for 3.5 h, then placed at 20 °C for an additional 26 h. Cells were pelleted and resuspended in 50 mM Tris (pH 8.0), 500 mM NaCl, 20 mM imidazole with 1 μM pepstatin, 100 μM PMSF, and DNaseI (Sigma, Kanagawa, Japan). The cells were lysed in a French pressure cell and centrifuged at 40,000× g for 20 min at 4 °C. Saturated ammonium sulfate was added to achieve a final solution of 10% and inverted several times. The lysate was centrifuged at 40,000× g for 10 min at 4 °C. The pellet was discarded, and the supernatant was increased to a final concentration of 50% ammonium sulfate. The supernatant was rotated for 10 min at 4 °C then centrifuged at 40,000× g for 10 min at 4 °C. The pellet was split into 150 mg aliquots. The WT h5-LOX used for the kinetics in this work was not purified due to a dramatic loss in activity and was therefore prepared as an ammonium-sulfate precipitate.

3.4. IC50 Determination

IC50 values for each acyl-CoA were measured against h5-LOX, h12-LOX, h15-LOX-1, and h15-LOX-2. The reactions were carried out in 25 mM HEPES buffer (pH 8.00), 0.01% Triton X-100, and 10 μM AA for h12-LOX; 25 mM HEPES buffer (pH 7.5), 0.01% Triton X-100, and 10 μM AA for h15-LOX-1 and h15-LOX-2; and 50 mM HEPES buffer (pH 7.5), 100 µM EDTA, 50 mM NaCl, 200 µM ATP, and 10 μM AA for h5-LOX. IC50 reactions were implemented with h15-LOX-1 (0.125 μM), h12-LOX (0.300 μM), h15-LOX-2 (0.5 μM), and h5-LOX (∼600 nM ammonium sulfate salt) at 22 °C in a 1 cm2 quartz cuvette containing 2 mL of the according buffer and 10 μM AA for no longer than 2 min to avoid any cleavage of the acyl-coenzyme A. Inhibitor concentrations ranging from 50 μM to 0.01 μM were used. All reactions were conducted using a Cary-UV Vis spectrophotometer (Agilent, Santa Clara, CA, USA). The fastest rates over 15 s intervals were recorded. IC50 values were obtained by determining the percent of inhibition of the enzymatic rate at ten inhibitor concentrations and plotting the rate against the inhibitor concentration, followed by a hyperbolic saturation curve fit. The saturation curve fits were performed in duplicate or triplicate, depending on the quality of the data. It should be noted that stability measurements of the acyl-CoAs indicated that there was no appreciable hydrolysis in the short timeframe of the kinetic measurements (approximately one minute). Through HPLC monitoring, we were able to determine a lack of degradation of the fatty acyl-CoAs over the first hour.

3.5. Enzymatic Product Determination

h15-LOX-2 (0.200 μM), h15-LOX-1 (0.125 μM), h5-LOX (0.300 μM), and h12-LOX (0.300 μM) were reacted with 30 μM of acyl-CoAs in 2 mL of 25 mM HEPES (pH 7.5 for h15-LOX-1 and h15-LOX-2, pH 8.0 for h12-LOX) at room temperature with ambient oxygen. The same buffer as used to determine IC50 values for h5-LOX was used for product determination. Turnover was monitored by absorbance at 234 nm.

3.6. Kinetic Investigation of Oleoyl-CoA (18:1) Inhibition of h15-LOX-2

The reactions were carried out in 25 mM HEPES buffer (pH 7.5), 0.01% Triton X-100. Varying amounts of substrates from AA ranging from 50 μM to 0.01 μM and 0, 150, 300, 450, and 600 nM of 18:2 acyl-coA were used. Reactions were performed at 22 °C in a 1 cm2 quartz cuvette containing 2 mL of the according buffer for no longer than 2 min to avoid any cleavage of the acyl-coenzyme A. Using a Cary-UV Vis spectrophotometer, the fastest rates over 15 s intervals were recorded. Kinetic values were obtained by plotting the rate against the substrate concentration, followed by a hyperbolic saturation curve fit. The saturation curve fits were performed in triplicate and on separate days to ensure there was no bias. The equations used to find kinetic values were from previous studies [32].

4. Conclusions

In conclusion, our previously reported cryo-EM structure for h12-LOX indicated an acyl-CoA bound to its active site, most likely oleoyl-CoA [30]. Based on this discovery, we investigated the inhibitory potency of acyl-CoAs against lipoxygenase isozymes and were able to determine that acyl-CoAs can be micromolar inhibitors against LOX isozymes, with oleoyl-CoA (18:1) being the most potent, having nanomolar allosteric potency against h15-LOX-2. Surprisingly, LA-CoA is a rapid substrate for h15-LOX-1, suggesting its product could have biological activity. Considering that acyl-CoAs exist in the cell at micromolar concentrations [38], it is reasonable to assume that acyl-CoAs could have a significant effect on the cellular activity of LOX isozymes, both as inhibitors and substrates. We are currently investigating whether the cellular concentrations of acyl-CoAs affect LOX activity in vivo and if this activity affects LOX signaling.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241310941/s1.

Author Contributions

Conceptualization, A.G., M.H. and T.H.; methodology, M.T. and K.Y.; investigation, M.T.; writing—original draft preparation, M.T., K.Y. and T.H.; writing—review and editing, M.T., K.Y., A.G., M.H. and T.H.; supervision, M.T. and T.H.; project administration, T.H.; funding acquisition, M.H. and T.H. All authors have read and agreed to the published version of the manuscript.

Funding

NIH-GM131835.

Data Availability Statement

No archived data for this work. All data presented.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

LOX, lipoxygenase; h12-LOX, human platelet 12S-LOX; h15-LOX-1, human reticulocyte 15S-LOX-1; WT h5-LOX, human 5S-LOX; h15-LOX-2, human epithelial 15S-LOX-2; r15-LOX-1, r12/15-LOX, or rALOX15, rabbit 15S-LOX-1; c11-LOX, coral 11R-LOX; SLO-1, soybean lipoxygense-1; AA, arachidonic acid; DHA, docosahexaenoic acid; 12(S)-HpETE, 12(S)-hydroperoxyeicosatetraenoic acid; 12(S)-HETE, 12(S)-hydroxyeicosatetraenoic acid; COX, cyclooxygenase; ML355, h12-LOX specific inhibitor; NSAIDs, nonsteroidal anti-inflammatory drugs; coxib, COX-2 selective inhibitor; ICP-MS, inductively coupled plasma mass spectroscopy; SEC, size exclusion chromatography; BSA, bovine serum albumin; DCM, dichloromethane; DMF, dimethylformamide; SDS-PAGE, sodium dodecyl sulfate-polyacrylamide gel electrophoresis; EDTA, ethylenediaminetetraacetic acid; HEPES, N-2-hydroxyethylpiperazine-N’-2-ethanesulfonic acid; PMSF, phenylmethylsulfonyl fluoride; Palmitoyl-CoA (16:0), palmitoyl-coenzyme A (16:0); Palmitoleoyl-CoA (16:1), palmitoleoyl-coenzyme A (16:1); Stearoyl-CoA (18:0), stearoyl-coenzyme A (18:0); Oleoyl-CoA (18:1), oleoyl-coenzyme A (18:1); Linoleoyl-CoA (18:2), linolenoyl-coenzyme A (18:2); ɣ-Linolenoyl-CoA (18:3), ɣ-linolenoyl-coenzyme A (18:3); Arachidonoyl-CoA (20:4), arachidonoyl-coenzyme A (20:4); Docosahexaenoyl-CoA (22:6), docosahexaenoyl-coenzyme A (22:6); Methyl-13Z,16Z-docosadienoic, methyl-13Z,16Z-docosadienoic (22:2); Methyl-13(Z),16(Z),19(Z)-Docosatrienoate, methyl-13(Z),16(Z),19(Z)-docosatrienoate (22:3); Methyl-9(Z),13(Z),16(Z),19(Z)-docosatetraenoate, methyl-9(Z),13(Z),16(Z),19(Z)-docosatetraenoate (22:4).

References

  1. Feingold, K.R. The Effect of Endocrine Disorders on Lipids and Lipoproteins. In Endotext; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Dungan, K., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  2. Horn, A.; Jaiswal, J.K. Structural and signaling role of lipids in plasma membrane repair. Curr. Top Membr. 2019, 84, 67–98. [Google Scholar] [PubMed]
  3. Markovic, M.; Ben-Shabat, S.; Aponick, A.; Zimmermann, E.M.; Dahan, A. Lipids and Lipid-Processing Pathways in Drug Delivery and Therapeutics. Int. J. Mol. Sci. 2020, 21, 3248. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, D.; Wei, Y.; Huang, Q.; Chen, Y.; Zeng, K.; Yang, W.; Chen, J.; Chen, J. Important Hormones Regulating Lipid Metabolism. Molecules 2022, 27, 7052. [Google Scholar] [CrossRef]
  5. Reddy, P.; Jialal, I. Biochemistry, Fat Soluble Vitamins. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  6. Feingold, K.R. Obesity and Dyslipidemia. In Endotext; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Dungan, K., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  7. Singla, P.; Bardoloi, A.; Parkash, A.A. Metabolic effects of obesity: A review. World J. Diabetes 2010, 1, 76–88. [Google Scholar] [CrossRef]
  8. Parhofer, K.G. Interaction between Glucose and Lipid Metabolism: More than Diabetic Dyslipidemia. Diabetes Metab. J. 2015, 39, 353–362. [Google Scholar] [CrossRef] [Green Version]
  9. Athyros, V.G.; Doumas, M.; Imprialos, K.P.; Stavropoulos, K.; Georgianou, E.; Katsimardou, A.; Karagiannis, A. Diabetes and lipid metabolism. Hormones (Athens) 2018, 17, 61–67. [Google Scholar] [CrossRef] [Green Version]
  10. Bhargava, S.; de la Puente-Secades, S.; Schurgers, L.; Jankowski, J. Lipids and lipoproteins in cardiovascular diseases: A classification. Trends Endocrinol. Metab. 2022, 33, 409–423. [Google Scholar] [CrossRef] [PubMed]
  11. Yan, A.; Xie, G.; Ding, X.; Wang, Y.; Guo, L. Effects of Lipid Overload on Heart in Metabolic Diseases. Horm. Metab. Res. 2021, 53, 771–778. [Google Scholar] [CrossRef]
  12. Craig, M.; Yarrarapu, S.N.S.; Dimri, M. Biochemistry, Cholesterol. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  13. Ahmed, S.; Shah, P.; Ahmed, O. Biochemistry, Lipids. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  14. Wakil, S.J.; Abu-Elheiga, L.A. Fatty acid metabolism: Target for metabolic syndrome. J. Lipid Res. 2009, 50, S138–S143. [Google Scholar] [CrossRef] [Green Version]
  15. Kelly, D.J.; Hughes, N.J. The Citric Acid Cycle and Fatty Acid Biosynthesis. In Helicobacter Pylori: Physiology and Genetics; Mobley, H.L.T., Mendz, G.L., Hazell, S.L., Eds.; ASM Press: Washington, DC, USA, 2001. [Google Scholar]
  16. Haddad, A.; Mohiuddin, S.S. Biochemistry, Citric Acid Cycle. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  17. Shi, L.; Tu, B.P. Acetyl-CoA and the regulation of metabolism: Mechanisms and consequences. Curr. Opin. Cell Biol. 2015, 33, 125–131. [Google Scholar] [CrossRef] [Green Version]
  18. Martinez-Reyes, I.; Chandel, N.S. Acetyl-CoA-directed gene transcription in cancer cells. Genes Dev. 2018, 32, 463–465. [Google Scholar] [CrossRef] [PubMed]
  19. Felix, J.B.; Cox, A.R.; Hartig, S.M. Acetyl-CoA and Metabolite Fluxes Regulate White Adipose Tissue Expansion. Trends Endocrinol. Metab. 2021, 32, 320–332. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, Y.; Yu, W.; Li, S.; Guo, D.; He, J.; Wang, Y. Acetyl-CoA Carboxylases and Diseases. Front. Oncol. 2022, 12, 836058. [Google Scholar] [CrossRef] [PubMed]
  21. Maloberti, P.M.; Duarte, A.B.; Orlando, U.D.; Pasqualini, M.E.; Solano, A.R.; Lopez-Otin, C.; Podesta, E.J. Functional interaction between acyl-CoA synthetase 4, lipooxygenases and cyclooxygenase-2 in the aggressive phenotype of breast cancer cells. PLoS ONE 2010, 5, e15540. [Google Scholar] [CrossRef] [Green Version]
  22. Sakuma, S.; Fujimoto, Y.; Katoh, Y.; Kitao, A.; Fujita, T. The effects of fatty acyl CoA esters on the formation of prostaglandin and arachidonoyl-CoA formed from arachidonic acid in rabbit kidney medulla microsomes. Prostaglandins Leukot Essent Fat. Acids 2001, 64, 61–65. [Google Scholar] [CrossRef]
  23. Bannenberg, G.; Serhan, C.N. Specialized pro-resolving lipid mediators in the inflammatory response: An update. Biochim. Biophys. Acta 2010, 1801, 1260–1273. [Google Scholar] [CrossRef] [Green Version]
  24. Basil, M.C.; Levy, B.D. Specialized pro-resolving mediators: Endogenous regulators of infection and inflammation. Nat. Rev. Immunol. 2016, 16, 51–67. [Google Scholar] [CrossRef]
  25. Yang, A.; Wu, Y.; Yu, G.; Wang, H. Role of specialized pro-resolving lipid mediators in pulmonary inflammation diseases: Mechanisms and development. Respir. Res. 2021, 22, 204. [Google Scholar] [CrossRef]
  26. Schebb, N.H.; Kuhn, H.; Kahnt, A.S.; Rund, K.M.; O’Donnell, V.B.; Flamand, N.; Peters-Golden, M.; Jakobsson, P.J.; Weylandt, K.H.; Rohwer, N.; et al. Formation, Signaling and Occurrence of Specialized Pro-Resolving Lipid Mediators-What is the Evidence so far? Front. Pharmacol. 2022, 13, 838782. [Google Scholar] [CrossRef]
  27. Haeggstrom, J.Z.; Funk, C.D. Lipoxygenase and leukotriene pathways: Biochemistry, biology, and roles in disease. Chem. Rev. 2011, 111, 5866–5898. [Google Scholar] [CrossRef]
  28. Mashima, R.; Okuyama, T. The role of lipoxygenases in pathophysiology; new insights and future perspectives. Redox Biol. 2015, 6, 297–310. [Google Scholar] [CrossRef] [Green Version]
  29. Kuhn, H.; Banthiya, S.; van Leyen, K. Mammalian lipoxygenases and their biological relevance. Biochim. Biophys. Acta 2015, 1851, 308–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Mobbs, J.I.; Black, K.A.; Tran, M.; Venugopal, H.; Holman, T.R.; Holinstat, M.; Thal, D.M.; Glukhova, A. Cryo-EM structures of human arachidonate 12S-Lipoxygenase (12-LOX) bound to endogenous and exogenous inhibitors. bioRxiv 2023. [Google Scholar] [CrossRef]
  31. Segraves, E.N.; Holman, T.R. Kinetic investigations of the rate-limiting set in human 12- and 15-lipoxygenase. Biochemistry 2003, 42, 5236–5243. [Google Scholar] [CrossRef]
  32. Joshi, N.; Hoobler, E.K.; Perry, S.; Diaz, G.; Fox, B.; Holman, T.R. Kinetic and structural investigations into the allosteric and pH effect on the substrate specificity of human epithelial 15-lipoxygenase-2. Biochemistry 2013, 52, 8026–8035. [Google Scholar] [CrossRef] [Green Version]
  33. Aleem, A.M.; Tsai, W.C.; Tena, J.; Alvarez, G.; Deschamps, J.; Kalyanaraman, C.; Jacobson, M.P.; Holman, T. Probing the Electrostatic and Steric Requirements for Substrate Binding in Human Platelet-Type 12-Lipoxygenase. Biochemistry 2019, 58, 848–857. [Google Scholar] [CrossRef]
  34. Bishop, J.E.; Hajra, A.K. A method for the chemical synthesis of 14C-labeled fatty acyl coenzyme A’s of high specific activity. Anal. Biochem. 1980, 106, 344–350. [Google Scholar] [CrossRef] [Green Version]
  35. Blecher, M. Synthesis of long-chain fatty acyl-coA thioesters using N-hydroxysuccinimide esters. Methods Enzymol. 1981, 72, 404–408. [Google Scholar]
  36. Amagata, T.; Whitman, S.; Johnson, T.A.; Stessman, C.C.; Loo, C.P.; Lobkovsky, E.; Clardy, J.; Crews, P.; Holman, T.R. Exploring sponge-derived terpenoids for their potency and selectivity against 12-human, 15-human, and 15-soybean lipoxygenases. J. Nat. Prod. 2003, 66, 230–235. [Google Scholar] [CrossRef]
  37. Vasquez-Martinez, Y.; Ohri, R.V.; Kenyon, V.; Holman, T.R.; Sepulveda-Boza, S. Structure-activity relationship studies of flavonoids as potent inhibitors of human platelet 12-hLO, reticulocyte 15-hLO-1, and prostate epithelial 15-hLO-2. Bioorg. Med. Chem. 2007, 15, 7408–7425. [Google Scholar] [CrossRef] [Green Version]
  38. Ronowska, A.; Szutowicz, A.; Bielarczyk, H.; Gul-Hinc, S.; Klimaszewska-Lata, J.; Dys, A.; Zysk, M.; Jankowska-Kulawy, A. The Regulatory Effects of Acetyl-CoA Distribution in the Healthy and Diseased Brain. Front. Cell Neurosci. 2018, 12, 169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Equations were determined from Scheme 1 to determine Ki, α (the change in KM), β (the change in kcat), and β/α (the change in kcat/KM).
Scheme 1. Equations were determined from Scheme 1 to determine Ki, α (the change in KM), β (the change in kcat), and β/α (the change in kcat/KM).
Ijms 24 10941 sch001
Figure 1. KM (app) vs. concentrations of oleyl-CoA (18:1).
Figure 1. KM (app) vs. concentrations of oleyl-CoA (18:1).
Ijms 24 10941 g001
Figure 2. kcat/KM (app) vs. concentrations of oleyl-CoA (18:1).
Figure 2. kcat/KM (app) vs. concentrations of oleyl-CoA (18:1).
Ijms 24 10941 g002
Figure 3. kcat (app) vs. concentrations of oleyl-CoA (18:1).
Figure 3. kcat (app) vs. concentrations of oleyl-CoA (18:1).
Ijms 24 10941 g003
Scheme 2. Synthesis of DHA-CoA.
Scheme 2. Synthesis of DHA-CoA.
Ijms 24 10941 sch002
Table 1. IC50 values with lipoxygenase isozymes #.
Table 1. IC50 values with lipoxygenase isozymes #.
Acyl-Coenzyme Ah12-LOX * (µM) h15-LOX-1
(µM)
h15-LOX-2
(µM)
h5-LOX
(µM)
Palmitoyl-CoA (16:0)>200>500>2003.3 ± 0.3
Palmitoleoyl-CoA (16:1)>500>500>2002.0 ± 0.4
Stearoyl-CoA (18:0)>2004.2 ± 0.67.6 ± 1>50
Oleoyl-CoA (18:1)32 ± 439 ± 20.62 ± 0.060>50
Linoleoyl-CoA (18:2)>200(Substrate)>100>100
ɣ-Linolenoyl-CoA (18:3)>200>500>500>500
Arachidonoyl-CoA (20:4)110 ± 20>500>500>200
Docosahexaenoyl CoA (22:6)>500>500>500>200
# See the Supplementary Materials for data analysis (Figures S1–S8). * The h12-LOX data have been published previously [30].
Table 2. Kinetic parameters for h15-LOX-1 with linoleoyl-CoA.
Table 2. Kinetic parameters for h15-LOX-1 with linoleoyl-CoA.
kcat7.5 (0.4) s−1
KM12 (0.8) µM
kcat/KM0.62 (0.1) µM−1 s−1
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

Tran, M.; Yang, K.; Glukhova, A.; Holinstat, M.; Holman, T. Inhibitory Investigations of Acyl-CoA Derivatives against Human Lipoxygenase Isozymes. Int. J. Mol. Sci. 2023, 24, 10941. https://doi.org/10.3390/ijms241310941

AMA Style

Tran M, Yang K, Glukhova A, Holinstat M, Holman T. Inhibitory Investigations of Acyl-CoA Derivatives against Human Lipoxygenase Isozymes. International Journal of Molecular Sciences. 2023; 24(13):10941. https://doi.org/10.3390/ijms241310941

Chicago/Turabian Style

Tran, Michelle, Kevin Yang, Alisa Glukhova, Michael Holinstat, and Theodore Holman. 2023. "Inhibitory Investigations of Acyl-CoA Derivatives against Human Lipoxygenase Isozymes" International Journal of Molecular Sciences 24, no. 13: 10941. https://doi.org/10.3390/ijms241310941

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