key: cord-0803241-xb0g837o authors: Khatua, Biswajit; El-Kurdi, Bara; Patel, Krutika; Rood, Christopher; Noel, Pawan; Crowell, Michael; Yaron, Jordan R.; Kostenko, Sergiy; Guerra, Andre; Faigel, Douglas O.; Lowe, Mark; Singh, Vijay P. title: Adipose saturation reduces lipotoxic systemic inflammation and explains the obesity paradox date: 2021-01-29 journal: Sci Adv DOI: 10.1126/sciadv.abd6449 sha: c43ba530dcb21f2b4a01baef95c73289d3b5f8e8 doc_id: 803241 cord_uid: xb0g837o Obesity sometimes seems protective in disease. This obesity paradox is predominantly described in reports from the Western Hemisphere during acute illnesses. Since adipose triglyceride composition corresponds to long-term dietary patterns, we performed a meta-analysis modeling the effect of obesity on severity of acute pancreatitis, in the context of dietary patterns of the countries from which the studies originated. Increased severity was noted in leaner populations with a higher proportion of unsaturated fat intake. In mice, greater hydrolysis of unsaturated visceral triglyceride caused worse organ failure during pancreatitis, even when the mice were leaner than those having saturated triglyceride. Saturation interfered with triglyceride’s interaction and lipolysis by pancreatic triglyceride lipase, which mediates organ failure. Unsaturation increased fatty acid monomers in vivo and aqueous media, resulting in greater lipotoxic cellular responses and organ failure. Therefore, visceral triglyceride saturation reduces the ensuing lipotoxicity despite higher adiposity, thus explaining the obesity paradox. While quantitative parameters such as body mass index (BMI) (1), waist circumference, and amount of visceral adipose (2, 3) are well-studied risk factors for acute disease severity, the impact of adipose composition on severity is unclear. Organ failure, the hallmark of acute pancreatitis (AP) severity, has been associated with long-chain nonesterified fatty acid (NEFA) lipotoxicity in humans (4) (5) (6) (7) . Clues to the role of adipose composition in severity come from studies from Western populations showing a BMI of >30 mostly, but not always (2, 3) associated with severity, while in Eastern populations, much lower BMIs, sometimes ≥23 (8, 9) , are associated with severity. This "obesity paradox" (10) also occurs in other acute scenarios such as burns (11) , acute heart failure (12) , after trauma (13) , cardiovascular surgery (14) , and during critical illnesses (15) in which elevated pancreatic enzymes (16) (17) (18) and long-chain NEFA (19, 20) have been associated with severity. More recent studies associate such an enzyme leak (21) and dietary fat composition (22) , along with elevated NEFA levels, to the severity (23) and mortality from coronavirus disease 2019 (COVID-19) (22, 24) . However, the mechanisms determining NEFA generation and how NEFA mediate outcomes such as the cytokine storm (24) and organ failure (19, 20, 22) are unclear. Most individual studies supporting the obesity paradox come from the Western Hemisphere (12) (13) (14) , where dietary and adipose fat are more saturated (25, 26) . We therefore aimed at understanding this problem in different populations, by modeling AP, wherein the pancreatic lipases leak into the surrounding fat (6, 7, 27) and hydrolyze the neutral lipids (7) . Principal among these lipases is pancreatic triglyceride lipase (PNLIP), which enters visceral adipocytes by various mechanisms (7) and mediates lipotoxic systemic injury along with organ failure (7, 28) via the generated long-chain NEFA (7, 28) inhibiting mitochondrial complexes I and V (6) . However, the impact of long-chain NEFA saturation on systemic inflammation is unknown. We therefore approached the obesity paradox by examining visceral fat composition as influenced by dietary fat composition (29) . Epidemiologically, all clinical reports up to 2017 associating AP severity with BMI were grouped by cutoff BMI, and the dietary fat composition representative of the country from which the study originated was studied. The dietary fat data were based on publicly available information archived by the Food and Agriculture Organization (FAO) and were classified as saturated or unsaturated based on the presence of double bonds in long-chain (>12 carbon) NEFA composing the dietary triglycerides (described in Methods). After analyzing the human metadata, we tested the plausibility of fat composition affecting the outcomes of pancreatitis by altering mouse visceral fat composition to replicate human visceral fat (25, 26, 29) . Unexpectedly, the human and experimental models cumulatively showed that the U.S. Food and Drug Administration (FDA)-recommended higher unsaturated fat intake (https://health.gov/dietaryguidelines/2015/ resources/2015-2020_Dietary_Guidelines.pdf; https://www.fda.gov/ files/food/published/Food-Labeling-Guide-%28PDF%29.pdf, appendix F) increased the risk of severe AP (SAP). We therefore investigated the mechanistic basis of this finding and noted that saturation of a triglyceride makes its interaction with PNLIP energetically unfavorable and reduces its lipolysis. Moreover, in general, unsaturated NEFAs achieved higher monomeric concentrations in vivo and aqueous systems, consequently eliciting stronger biological responses and causing worse injury and organ failure. These features may explain how obese populations with higher visceral fat saturation may sometimes be protected from severe disease. less (8, 9, (45) (46) (47) (48) (49) . Adult AP typically occurs after the third decade (50, 51) . Since the composition of visceral fat necrosed during AP may be influenced by the dietary fat composition over the preceding years (26) , we analyzed dietary fat composition of different countries shown in Fig. 1A . The per capita fat consumption was calculated by averaging the yearly data (1970 to 2011) for each country. Countries with a BMI cutoff of ≥30 had higher per capita saturated fat consumption ( Fig. 1B and fig. S1 ). While the amount of unsaturated fat consumption was the same ( fig. S3A ), unsaturated fat comprised a higher percentage of fat intake in the countries with reports having a cutoff BMI of ≤25 (pink countries; Fig. 1 , A and C). Overall, there was a moderate correlation between the percentage of patients with SAP and the percentage of unsaturated fat intake ( fig. S3B ). On meta-analysis (Fig. 1D) , a significantly increased risk of severity was noted for cutoff BMIs of ≤25 [pink shade, odds ratio (OR) 2.8, CI 1.3 to P = 0.008] and also BMI of >30 (blue and greens, OR 2.7, CI 1.8 to 3.8, P < 0.001). Publication bias was not detected on the basis of a Funnel plot ( fig. S1C ), an Egger's regression, or a Begg and Mazumdar rank correlation. There were no differences in age, sex distribution, or etiology of AP between the two groups ( fig. S1 ). While the effect of genes and comorbidities on %SAP cannot be commented on in these data, meta-regression showed that %unsaturated fatty acid (UFA) was able to explain 33% of the heterogeneity in the rate of SAP, and neither age, AP etiology, nor per capita gross domestic product (GDP) correlated with SAP. Furthermore, while per capita GDP of countries averaged for 5 years preceding publication was significantly lower in countries reporting SAP at BMIs of ≤25 ($15,768 ± 14,624 versus $32,272 ± 17,007, P = 0.018), neither mortality (5.2 ± 6.9 versus 6.7 ± 4.3%, P = 0.698) nor the proportion of patients with SAP (24 ± 12.5 versus 20.7 ± 8.4%, P = 0.892) was different in the two BMI cutoff groups. Therefore, despite differences in income, the proportion of SAP and mortality were unchanged. This implied that quality of care was not different in the two groups of countries. Overall, these studies supported Studies using a BMI cutoff of ≤25 are shown in pink, those associating SAP with a BMI of ≥30 are shown in gold, and studies showing a BMI of >30 was not associated with severe AP (SAP) are show shown in blue. The respective countries from which the study originated are shown in pink, green, and blue, with a checkered pattern for countries with studies showing different outcomes. Box plot comparing the per capita per year saturated fat (from dairy, cattle, and palm oil) consumption in countries with BMI cutoffs of >30 and those with ≤25 BMI (B) and %UFA in dietary fat intake (C). (D) Meta-analysis showing OR for SAP for studies using a BMI cutoff of ≤25 (pink background) and those using a BMI cutoff of >30 (blue-yellow background). The overall OR is shown at the bottom (orange background). that severe pancreatitis occurred at a lower BMI in countries with a lower dietary saturated fatty acid (SFA) intake. Since we are not aware of the literature on how diet would affect the genetic background of a population over several generations to alter the severity of pancreatitis, we therefore went on to experimentally study whether and how altering visceral fat composition affected the development of SAP in the commonly used caerulein (CER) model of AP, which, in lean C57BL6 normal chow-fed mice, remains mild, self-limited, and without organ failure (27) . Dietary unsaturated fat results in unsaturated visceral fat and worsens AP more than higher amounts of saturated visceral fat We first aimed at recapitulating the human studies showing that dietary fat composition can change adipose tissue composition in an animal model (26) . Linoleic acid (LA; C18:2) is an unsaturated essential fatty acid (i.e., only available through diet) present at high concentrations in common FDA-recommended dietary fats (52) (e.g., cooking oils) and comprises a high proportion in pancreatic fat necrosis (FN) (5, 6) . Being essential, LA's concentrations in visceral fat parallel dietary intake (29) . We thus formulated a diet enriched in LA (70% LA in a 45% fat diet) to represent high %UFA intake in dietary fat (red box in fig. S4A and Fig. 2A1 , red columns) (26, 52) . Similarly, a diet enriched in saturated fat was formulated, which had palmitic acid (PA; C16:0, PA 68% of a 47% fat diet; fig. S4B , green rectangle) replicating the 65 to 70% saturation of dairy fat. These diets spanned the range of %UFA intake in humans (16% in Australia and 79% in Japan), as shown in fig. S3A . Mice fed these diets had a similar food intake (6.5 ± 1.4 g/day of UFA diet or 6.3 ± 0.5 g/day of SFA diet). This altered their visceral triglyceride composition, with LA and PA increasing to 40 to 45% ( Fig. 2A1 and detailed in fig. S5 ) in the UFA-and SFA-fed groups, respectively. We previously showed that a normal chow diet with 5% fat diet (Purina 5053), which contains ≥70% UFA and <20% SFA Table comparing the fatty acid composition of fat pad triglycerides (TG) and diets of mice given the SFA-and UFA-enriched diets. Body weights (A2), body fat (A3), and body fat as a percentage of body weight (A4) in the SFA-and UFA-fed mice. Serum amylase (B1) and lipase (B2) in control (CON) mice and after 24 hours of AP (CER). Local pancreatic injury seen histologically (C1) and quantified as acinar necrosis (C2) as a percentage of total parenchymal area and percentage of acinar necrosis adjacent to the FN (shown in yellow rectangles), termed as "% peri-fat acinar necrosis" (C3). Systemic injury measured as renal injury [TUNEL staining showing brown nuclei in (D1) and serum BUN in (D2)], lung TUNEL positivity highlighted with arrows in (D3), and shown as number per high-power field (HPF) in (D4), along with shock (as measured by a drop in carotid pulse distention) in (E1), serum calcium (E2), and survival curve (E3) of mice with CER AP. NS, not significant. consistent with FDA recommendations, results in fat pad triglyceride to be composed of 18 ± 6% PA and 31 ± 14% LA, with 35 ± 6% oleic acid (OA), i.e., C18:1 in ob/ob mice (7). On feeding the special diets, the LA concentrations in the fat pads of the UFAand SFA-fed ob/ob mice correspondingly matched reports of those in human adipose tissue from Japan [≈40% (25) ] and the United States [5% (26) ]. By 8 to 14 weeks, the mice averaged 45.5 ± 0.5 g in both groups. CER AP reduced survival to 10% in the UFA group by day 3 versus 90% in the SFA group (P < 0.02; fig. S6 ). We then simulated the lower cutoff BMI (≤25) associated with SAP in countries with a higher %UFA intake. AP was initiated in UFA-fed mice weighing 20 to 30% less and having 35 to 40% less adipose tissue (Fig. 2, A2 to A4). AP increased serum amylase and lipase similarly in both groups (Fig. 2, B1 and B2) but was worse in the leaner UFA group. The SFA group has lesser pancreatic necrosis (5.8 ± 0.8 versus 16.8 ± 4.7%, P = 0.024; Fig. 2 , C1 and C2) especially bordering FN, termed peri-fat acinar necrosis (2.7 ± 0.4 versus 6.9 ± 1.6%, P = 0.03; green outline, Fig. 2 , C1 and C3). Consistent with SAP (53, 54) , the UFA-fed mice had greater lung and renal tubular TUNEL (terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate nick end labeling) positivity, higher serum blood urea nitrogen (BUN) levels ( Fig. 2, D1 to D4), a greater decrease in carotid pulse distention (consistent with shock), and SAP-associated hypocalcemia (55, 56) (Fig. 2, E1 and E2), resulting in 20% survival (P < 0.02 versus 90% in the SFA group; Fig. 2E3 ). The findings of SAP were replicated in UFA-fed C57BL6 mice with diet-induced obesity (DIO), but not the normal chow-fed mice (30.8 ± 0.7 g), which had mild pancreatitis ( fig. S7 , A to F) (27) . However, we did not use DIO as the primary model since the time to weight gain (>5 months) was prolonged. It has recently been shown that PNLIP mediates FN and severity of AP (7) . To understand the worse outcomes in UFA-fed mice, we compared the necrosed gonadal fat pads in both groups (Fig. 3A) . Grossly, the SFA group has reduced FN; however, the pancreatic amylase, lipase activity, and PNLIP protein (6-to 14-fold control) had a large increase in both groups ( The UFA mice had a larger cytokine mRNA increase in the fat pads during AP and interleukin-6 (IL-6), monocyte chemoattractant protein-1 (MCP-1), and tumor necrosis factor- (TNF-) proteins in the sera (Fig. 3 , C1 to C4), which correlated with lower IB- (42 ± 15% of controls, P < 0.05) in the necrosed UFA fat pads (Fig. 3B2 ). Since NEFA mediates cytokine increase (7), we measured serum NEFA to understand the underlying mechanism. The UFA group with AP had higher serum NEFA increase, especially UFAs, including C18:1 (OA) and C18:2 (LA) (Fig. 3, D1 to D4). This pattern has been noted during SAP-induced organ failure (7, 28, 53, 54) in both rodents (5, 6, 27) and humans (4) . However, inexplicably, the increase in C18:1, C18:2, and C16:1 (Fig. 3 , D1, D2, and D6) was muted in the SFA-fed mice, despite C18:1 being equal and C16:1 being higher in the fat pads of the SFA group ( Fig. 2A1 and fig. S5 ). Similarly, C16:0 and SFA overall (Fig. 3, D5 and D7, and fig. S8B ) increased more in the UFA group, despite the SFA group having more of these in the visceral triglyceride, serum at baseline. The proportion of SFA in general went down with AP (Fig. 3D8 ). This generalized reduction in SFA release suggested that SFAs in triglyceride are an unfavorable substrate for PNLIP. We next directly investigated whether visceral triglyceride composition, irrespective of obesity or genetic background, affects its lipolysis and consequent severity during AP. For this experiment, the triglycerides of LA (C18:2), glyceryl trilinoleate (GTL), or PA (C16:0), i.e., glyceryl tripalmitate (GTP), were intraperitoneally given to lean mice with a genetically dissimilar background than the C57Bl6 mice, i.e., lean CD-1 mice. Two mechanistically distinct AP models-i.e., CER and IL12/18, both of which are mild in lean rodents (27, 57)-were then induced ( Fig. 4 and fig. S9 ). Mice given GTL or GTP alone exhibited normal behavior, with no change in physiologic or biochemical parameters. During AP, neither GTL nor GTP affected the initial serum amylase or lipase increase (Fig. 4 , A1 and A2, and fig. S9 , A1 and A2), which is similar 18) , with GTL (IL12,18 + GTL) or GTP (IL12,18 + GTP). Note similar amylase and lipase at day 2 in all groups and progressive worsening in the GTL group, which has reduced survival. (B1 to B3) Representative H&E-stained images of the pancreas of mice belonging to the groups mentioned below the image. Note the increased necrosis at the periphery of the lobules in IL12,18 + GTL group (black arrows). Box plots showing serum glycerol (C1) and NEFA (C2) concentrations in control (CON) mice and in the different pancreatitis groups. (D1 to D3) Representative TUNEL-stained images of the kidneys of mice belonging to the groups mentioned below the image. Note the increased positivity in the tubules of the IL12,18 + GTL group (black arrows), which also has a significantly higher BUN than other groups on ANOVA (*) (D4). Box plots of the serum resistin (E1), IL-6(E2), TNF- (E3), and MCP-1(E4) control (CON) mice and in the different pancreatitis groups. * indicates significantly higher than control on ANOVA, and # indicates a significant reduction compared to the IL12,18 + GTL group. to ob/ob mice given UFA or SFA diets (Fig. 2, B1 and B2). The AP outcomes also paralleled the UFA and SFA groups (Figs. 2 and 3) with 0% survival in the GTL groups versus 80 to 90% in the GTP groups (P < 0.01; Fig. 4A3 and fig. S9A3 ). The GTL group had worse pancreatic necrosis (Fig. 4 , B1 to B3, and fig. S9, B1 and B2), predominantly at the periphery of the lobules exposed to the lipolytically generated LA, resembling peri-fat acinar necrosis (Fig. 2, C1 to C3). The higher serum glycerol and corresponding NEFA in the GTL groups with AP ( Fig (7), serum IL-6, TNF-, and MCP-1 were higher in both AP models with GTL ( Fig. 4 , E2 to E4, and fig. S9, E2 to E4), perhaps due to the systemic lipotoxicity of LA noted as higher serum damage-associated molecular patterns (DAMPs), i.e., doublestranded DNA (dsDNA) and histone-complexed DNA fragments in the GTL groups with pancreatitis (fig. S11, C, D, F, and G). These findings suggested that unsaturated visceral triglyceride is hydrolyzed more than saturated triglyceride and worsens systemic inflammation and organ failure, consistent with what we note epidemiologically ( Fig. 1) . We thus went on to study the mechanistic basis of this in more detail. To understand how visceral triglyceride composition influences AP severity, we simulated the in vivo AP-associated lipase leak into fat using an in vitro system (Fig. 3 , B1 and B2) (6, 7, 27) . Acini release pancreatic enzymes into the surrounding medium under the basal state in vitro (58) . Thus, we measured the hydrolysis of triglycerides added to the medium and biological responses of the NEFA generated. Addition of pure triglycerides GTP, glyceryl trioleate (GTO; the triglyceride of OA, C18:1), or GTL (300 M each) to the stirred acinar medium in a quartz cuvette at 37°C resulted in 60 to 70% hydrolysis of GTL and GTO within 15 min but <10% of GTP (Fig. 5 , A1 and A2). Lipolysis of GTL and GTO, unlike GTP or GTLO (i.e., GTL + the lipase inhibitor orlistat at 50 M) which has <15% lipolysis, also caused acinar mitochondrial depolarization (m) and cytosolic calcium (Cai) increase (Fig. 5 , A3 and A4). Separately, on longer incubation (6 hours), 80 to 100% of GTL was hydrolyzed to glycerol (fig. S12A) versus 20 to 30% of GTP even at a 3× molar excess. GTL hydrolysis caused cytochrome c leakage, decreased ATP (adenosine 5′-triphosphate) levels, and increased LDH leakage and propidium iodide uptake (fig. S12, B to E). These studies showed that triglyceride saturation reduces its lipolysis and consequent biological effects. We then studied the effect of saturation on the hydrolysis of mixed triglycerides in the acinar medium, since triglycerides in vivo are mostly mixed. On the basis of pancreatic lipase lacking stereoselectivity for the sn-1/sn-3 position (59), we chose mixed triglycerides (each at 100 M) of LA, i.e., 1,2-dilinoleoyl-3-palmitoyl-rac-glycerol (LLP) (LA-LA-PA), 1,3-dipalmitoyl-2-linoleoylglycerol (PLP) (PA-LA-PA), and 1-palmitoyl-2-oleoyl-3-linoleoyl-rac-glycerol (LOP) (LA-OA-PA), and compared these to GTL (LA-LA-LA). Palmitate disproportionately reduced lipolysis over 15 min. For example, LLP generated <30% LA, and PLP generated <9% LA versus GTL (Fig. 5B1 ). Both linoleate (4.3 ± 1 M) and oleate (4.4 ± 1.2 M) generation were markedly reduced on LOP. This was paralleled by reductions in m and Cai increase (Fig. 5, B2 and B3). We then studied the interaction and hydrolysis of GTL, LLP, and LOP focusing solely on triglyceride hydrolysis by human PNLIP in phosphate-buffered saline (PBS; pH 7.4, 37°C, 150 mM Na). The hydrolysis, i.e., GTL > LLP > LOP, paralleled those in the acinar cell media (Fig. 5C1) . We thus studied their interaction with PNLIP using isothermal titration calorimetry (ITC). A single injection of human PNLIP (0.7 nmol/s) into a stable, sonicated, stirred, and 100 M suspension of these triglycerides in PBS (at 300 s; fig. S13A ) caused an endothermic interaction with a magnitude paralleling lipolysis, i.e., GTL > LLP > LOP, the enthalpy of which paralleled the raw heat data ( (60) and are more favorable than for LLP or LOP. Addition of orlistat (50 M) to PNLIP significantly reduced the H and V max in both types of injections (fig. S13, C, D, H, and I), supporting the relevance of the parameters to lipolysis. The above experiments thus showed that increasing saturation makes the interaction of unsaturated triglycerides with PNLIP energetically unfavorable, resulting in reduced lipolysis. To verify the experimental results in an independent unbiased manner, we undertook docking simulations using the open lid conformation of the human PNLIP-procolipase complex (1LPA) (61) and studied access of the relevant triglycerides into the PNLIP ligandbinding domain (LBD) via the oxyanion hole. Ser 152 , Asp 176 , and His 263 compose the catalytic triad of human PNLIP and are critical in fatty acid liberation from triglycerides. The closed conformation (1LBP) (62) was not used since our goal was not to study molecular dynamics of pancreatic lipase activation by interfaces. Previous in silico studies (63) found that orlistat, a potent lipase inhibitor, docks to pancreatic lipase with a GlideScore of −6.90 kcal/mol with root mean square deviation (RMSD) between 1.2 and 4.8 Å. This value served as a threshold to quantify strong binding. To verify this value, we docked orlistat to 1LPA with the induced fit protocol, and a similar GlideScore value was produced. Orlistat docked with a distance of 3.74 Å between the catalytic serine and the -lactone ring that inhibits PNLIP hydrolysis. The induced fit protocol docked GTL into the 1LPA LBD with a GlideScore of −7.16 and with a distance of 4.04 Å between the hydroxyl group of Ser 152 and the carbonyl C atom of the triglyceride's glycerol backbone (Fig. 5D1) . LLP docked with a GlideScore of −4.72 kcal/mol at a distance of 9.99 Å from Ser 152 , and LOP and GTP respectively produced GlideScores of −1.33 and −1.58 while being 12.42 and 11.98 Å from the catalytic serine (Fig. 5, D2 to D4). To verify the integrity of the docking simulation, the ligand present in the 1LPA crystal structure (61), dilauryl phosphatidyl choline (DLPC), was removed and then docked with the same induced fit docking protocol used for the three triglycerides. DLPC docked with a GlideScore of −7.46 (fig. S14A) and at a distance of 3.59 Å from the hydroxyl group of the catalytic serine, which is 0.39 Å from its location in the crystal structure (3.20 Å). The resolution of 1LPA is 3.04 Å, so the variance seen is within the margin of error. DLPC inhibited the lipolysis of GTL added simultaneously in a dose-dependent manner ( fig. S14B ). This inhibition by DLPC, along with the redocking closely recapitulating the crystallographic findings, thus validates the docking simulation. Overall, these studies show that long-chain SFAs like palmitate make a triglyceride's interaction with PNLIP structurally and energetically unfavorable, thus reducing its hydrolysis. We lastly went on to study how the fatty acids generated may result in cell injury and consequent organ failure. We first compared the ability of NEFA to directly induce inflammation and organ failure. LA [0.3% body weight (28, 53, 54) ], OA (7), or PA (0.5%) body weight was instilled into the peritoneal cavity to simulate NEFA generated from visceral fat (2 to 10% body weight) lipolysis (64) . Unlike LA and OA, there was no increase in circulating cytokines or organ failure (BUN increase) induced by PA (Fig. 6, A and B) . The median survival for LA-treated mice was 18 hours and for OA was 64 hours, while the PA group had no adverse outcome over 3 days. Serum NEFA (which are predominantly albumin bound) and unbound NEFA (uNEFA; Fig. 6 , C and D) increased significantly only in the LA and OA groups, along with levels of the DAMP and dsDNA (Fig. 6E ) and a drop in serum albumin consistent with the hypoalbuminemia noted during experimental (fig. S15, A to C) and clinical SAP (55) . This suggested that, unlike PA, both LA and OA with high uNEFA may directly injure cells, triggering DAMP release and downstream inflammation. This would also be consistent with the in vivo AP models (Figs. 3 and 4 and figs. S9 and S11, B to G) where higher NEFA were associated with worse outcomes. We thus compared the ability of the most prevalent long-chain NEFA (C18:2, C18:1, and C16:0; Fig. 6 , F1 to F3) to exist in a nonmicellar monomeric form in aqueous media. On ITC at 37°C, we noted the critical micellar concentration (CMC) and therefore the aqueous monomeric NEFA concentrations to increase with the number of double bonds (Fig. 6, F1 to F3 ). The CMCs of PA (C16:0), OA (C18:1), and LA (C18:2) were respectively <8, ≈40, and ≈160 M, which paralleled the uNEFA levels in vivo (Fig. 6D) . The calorimetric results were then validated using ultracentrifugation. Both of these methods could be performed at room temperature (23°C; fig. S16 ) and showed that the nonmicellar NEFA concentrations in the infranatents of 500 M NEFA in PBS (pH 7.4), spun at 10 5 g for 1 hour (fig. S16, A and B) , were similar to the CMCs on calorimetry (fig. S16, C and D) and those shown previously using diphenyl hexatriene fluorescence spectroscopy (65) . Thus, the CMC noted on calorimetrically at 37°C (Fig. 6, F1 to F3 ) are accurate. The role of uNEFA in causing biological effects was then studied in cells. Addition of 100 M LA (below CMC) to pancreatic acini caused a larger increase in Cai (Fig. 6G) and m (Fig. 6H ) than 100 M OA (above CMC). PA caused no change in these. A total of 0.5% albumin (Alb) completely aborted the Cai increase by both LA and OA and prevented m from progressing (Fig. 6, H and I) . This effect of albumin proves that the Cai and m increases were from the uNEFA. It also shows that Cai and m changes result from distinct mechanisms of these uNEFA. The sustained m elevation, without progression, despite albumin (Fig. 6 , H and I) signifying irreversible uNEFA toxicity, is distinct from the complete reversal of Cai by albumin (Fig. 6G ) and explains the cell death noted from LA or OA in previous studies (6) . Therefore, the greater Cai and m induced by LA than OA at 100 M (Fig. 6 , G and H) may be due to LA's higher CMC (Fig. 6F3) and thus explains the higher uNEFA levels and shorter survival noted in LA-treated mice. We tested this by measuring the dose response of increase in m over baseline (m) 60 s after the addition of LA or OA to acini (Fig. 6J ) or human embryonic kidney (HEK) 293 cells (Fig. 6K) to represent the pancreatic and kidney injury in vivo (Fig. 4, D1 to D4, and fig. S9 , D1 and D2). Consistent with OA's CMC (≈40 M; Fig. 6F3 ), m increased over baseline at 50 M OA (*) and then plateaued (Fig. 6 , J and K). Similarly, LA's m increased till 100 M and ceased at 200 M, consistent with its CMC of ≈160 M. While LA's m equaled OA's at 50 M, it was more than OA's (#) at concentrations of ≥100 M. LA and OA (60 M), unlike PA, also reduced endothelial barrier integrity (fig. S15, D and E) , potentially explaining the hypotension (Fig. 2E1 and fig. S10C ) hypothesized to be from vascular leak during SAP (66), along with explaining the hypoalbuminemia ( fig. S15 , A to C) (55) in severe AP, and the uNEFA increase noted (Fig. 6D) . Overall, these studies cumulatively validate that double bonds increase monomeric long-chain NEFA concentrations and signaling in an aqueous environment, thus enhancing their lipotoxicity. Here, we find that a higher proportion of dietary unsaturated fat can worsen AP outcomes at a lower adiposity than seen in individuals with a higher proportion of saturated fats in their diet. We show that the higher likelihood of SAP associates with a higher unsaturated triglyceride content in visceral fat. This occurs because the presence of SFAs in triglycerides makes the interaction of the substrate with PNLIP structurally and energetically unfavorable. Moreover, the unsaturated NEFA generated by lipolysis can exist as monomers (65) in aqueous media, unlike saturated NEFA, resulting in injurious signaling, lipotoxic inflammation, and organ failure. This can potentially explain why higher dietary UFAs may result in worse AP in leaner animals and humans with lower BMIs compared to the more obese ones who consume a diet with higher proportions of saturated fat (Fig. 7) , resulting in the obesity paradox. We note that the obesity paradox in pancreatitis (10) holds true when the data are grouped and analyzed by BMIs using the World Health Organization (WHO) cutoffs (67) relevant to the countries from which the study originated (Fig. 1) , including those that use BMI cutoffs of ≤25, which have a lower SFA and higher %UFA consumption in diet (Fig. 1, B and C) . Socioeconomic and quality-ofcare issues, age, sex, and etiology of AP are unlikely to have influenced our findings since the rate of SAP and mortality were the same in the >30 versus ≤25 BMI cutoff groups. We note that the SAP rates and %UFA intake have a moderate correlation ( fig. S3B) , and on meta-regression, %UFA intake explains 33% of the heterogeneity in SAP rates. Thus, other factors such as male sex, genetic background, and comorbidities that are not accounted for in the studies may contribute to the remaining heterogeneity in SAP rates. The difference in BMI cutoffs set by the WHO is commonly attributed to a higher percentage fat per body mass in Asian populations compared to Caucasian populations (68) . However, this is brought into question since visceral adiposity (3) that undergoes visceral FN during SAP (69) differs by ethnic groups (70) and in women versus men (71) . We note that all visceral fat is not the same, since a higher proportion of diet-derived UFAs like LA make smaller , left) or unsaturated triglyceride shown in red ( , right) is exposed to pancreatic lipase, principal among which is PNLIP (zoomed circles). Despite being lesser in amount, the more unsaturated triglyceride is hydrolyzed more into the lipotoxic NEFA ( ), which are stable as monomers at higher concentrations than saturated NEFA, that form micelles ( ) at lower concentrations. This results in worse systemic inflammation, injury, and organ failure in those with a higher unsaturated fat consumption, despite having less adiposity. This pathophysiology can explain the obesity paradox. amounts of visceral fat more prone to lipolysis (Figs. 2 and 3) , generating higher concentrations of monomeric NEFA (Fig. 6) , thus worsening outcomes in leaner populations (Fig. 1) . The reverse holds true for SFAs, and this is relevant to studies from Western countries (72) that show AP severity to be independent of intra-abdominal fat amounts even when this fat is above the range in studies from Asian countries (71) . This is exemplified by the six studies reporting a BMI of >30 to not be associated with severe AP-all of which came from countries with <40% UFA as dietary intake (Fig. 1C) . Our epidemiologic studies are limited by being retrospective, not involving dietary history, visceral fat sampling, or genetic studies, and by using nationwide dietary and economic data. We thus verified the conclusions in mice, cell, and pure enzyme models and physical chemistry. The use of two different strains of mice (i.e., C57bl/6 and CD-1) and two modes of increasing visceral triglyceride (i.e., obesity and intraperitoneal injection of triglyceride) experimentally support that acute triglyceride lipolysis is crucial in the pathogenesis of organ failure and weakens the argument in favor of the genetic or chronic effects of obesity such as insulin resistance and atherosclerosis in explaining these outcomes. The deleterious effects that we note with GTL are also seen with the triglyceride of OA but not with GTP during pancreatitis (5) . These along with the deleterious effect of monomeric unsaturated NEFA (Fig. 6) , lipase (16) (17) (18) , and NEFA elevation (19, 20, 23) , being associated with worse outcomes in other diseases states including COVID-19 (22, 24) , perhaps support the general relevance of dietary and visceral fat unsaturation in causing the obesity paradox (11) (12) (13) (14) (15) . A palmitate-enriched diet in ob/ob mice helped avoid the confounding effects of maldigestion of a dietary saturated triglyceride by pancreatic lipases, which we note as reduced lipolysis of GTP, LLP, and PLP (Fig. 5, A2 and B1 ). We used PNLIP since recent studies show that adipocyte triglyceride lipase, PNLIPRP2, and carboxyl ester lipase are unlikely to mediate lipotoxic systemic inflammation (7, 53) . Moreover, the similar lipolysis pattern in acinar media, which contains all three lipases (Fig. 5, A1 to A4 and B1 to B3) and PNLIP (Fig. 5, C1 to C3), lends credence of the concept to other lipases. Stearoyl-CoA (coenzyme A) desaturase-1(SCD-1) (73) is likely responsible for converting the dietary palmitate (C16:0) to palmitoleate (C16:1), which comprised 17% of the visceral triglyceride of SFA mice (Fig. 2A1 and fig. S5 ), despite palmitoleate being absent in the diet. SCD-1 is highly expressed in adipose tissue and can put a double bond in stearoyl-CoA or palmitoyl-CoA, thus forming oleate (C18:1) and palmitoleate from the palmitate in the diet. This may also explain the similar amounts of C18:1 noted in the visceral triglyceride of the UFA-and SFA-fed mice. Despite the potential influence of SCD-1 on our hypothesis, the similarity of dietary and visceral fat composition that we note in mice, the previous studies note in humans (26) , along with the lack of known pathways to saturate C18:1, and our mechanistic data cumulatively support our conclusions. An additional limitation may lie in our inability to resolve the lipolytic impact of stereochemical structure of the triglycerides (74), which we used in the racemic form rather than the enantiopure form. While this remains beyond our scope, the previous findings that pancreatic lipase hydrolyzes triglyceride without any stereoselectivity for the sn-1 or sn-3 position (59) suggest that the interference by palmitate is independent of these locations on the glycerol backbone. Last, the mechanisms underlying the exothermic changes (Fig. 5, C2 and C3) following the initial interaction of PNLIP and triglycerides remain unclear. The relative magnitude of these, i.e., GTL > LLP > LOP, mirror the initial interaction. Previous studies suggest that this may be due to binding of the generated LA to aromatic residues on PNLIP and further increasing its activity (75) . In summary, diet-induced visceral fat unsaturation increases lipolytic generation of unsaturated monomeric NEFA that cause cell injury, systemic inflammation, and organ failure. Long-chain SFAs like palmitate, however, interfere with this lipolysis, generating lower amounts of monomeric NEFA, making pancreatitis milder despite excess adiposity, thus explaining the obesity paradox (Fig. 7) . The FDA guidelines to reduce saturated and increase unsaturated fat intake (https://www.fda.gov/files/food/published/Food-Labeling-Guide-%28PDF%29.pdf, appendix F) (26, 52) could therefore potentially contribute to worsening systemic inflammation and organ failure. To explore the impact of BMI on AP severity, we conducted a literature review by searching the PubMed database from inception to January 2017). The terms "BMI" or "obese" and "acute pancreatitis" as well as "severe" or "severity" were used. A total of 118 studies were retrieved, as shown ( fig. S2 ). Seventy-nine studies were excluded, as they were not related to the subject of this review, as were 12 studies that were reviews/meta-analyses/hypotheses. The remaining studies were extensively reviewed for eligibility criteria, which were as follows: (i) clear BMI cutoff to define obesity, (ii) clear report of the incidence of mild AP (MAP) and SAP in obese versus nonobese groups, and (iii) clear and satisfactory definitions of MAP and SAP. Twenty-seven studies met our eligibility criteria and were therefore included in this review (fig. S1 ). These are shown in different colors in Fig. 1A for descriptive purposes. The studies were then categorized into those using a BMI of 30 as a cutoff (n = 20, the one with a cutoff of >29 was included here) and those that used a cutoff BMI of ≤25 (n = 7; with text and country in pink) to define obesity's association with SAP. The dietary fat consumption in the countries from which these papers originated was then extracted as described below and compared ( fig. S1 ). The incidence of MAP and SAP in both obese and nonobese patients was extracted for the purpose of meta-analysis. Total number of patients was divided into MAP and SAP and then into those that were obese versus nonobese. In cases where this was not explicitly clear, the communicating author was contacted (30, 45) , and the numbers so provided were included in the study. Papers that did not have adequate data or for reasons mentioned in the last column of fig. S1 and detailed in fig. S2 were excluded from the meta-analysis (n = 7). When more than one BMI cutoff was mentioned, the BMI used was the one at which the SAP risk increased. The proportion of SAP in all pancreatitis was calculated and compared between obese and nonobese groups. Subgroup analysis was performed on the basis of the definition of obesity used in the different studies included. Random effects model was used to calculate pooled OR with 95% CIs. Heterogeneity was assessed using the I 2 measure and the Cochran Q statistic. Statistical analysis was performed using Comprehensive Meta-Analysis Software version 3.3.070 (Biostat, Englewood, NJ, USA). Publication bias was assessed for using Begg and Mazumdar's test as well as Egger's regression intercept. The per capita per year dietary fat consumption of each country was collected from the FAO website (http://www.fao.org/faostat/en/#data/FBSH), which is reported as kilograms per person per year. For purposes of simplicity, the chief sources of long-chain saturated fat (dairy, red meat, and palm oil) were analyzed separately from those of unsaturated fat (fish and vegetable oil). Fat from dietary sources were calculated by percentage weight (3.25% for whole milk, 20% from red meat, and 10% for fish) unless the food source was directly fat (e.g., cream, butter, ghee, cheese, vegetable oil, or palm oil). Data from the country from which each paper originated were included in the categories "BMI >30 not related to SAP," "BMI >30 associated with SAP," and cutoff "BMI ≤25" (shown as blue, green with golden text or symbols, and pink in Fig. 1, respectively) . Each country was represented once in the category for a contributed paper (shown in Fig. 1, B and C) . Comparisons were done by grouping the BMI of >30 not related to SAP and BMI of >30 associated with SAP into a grouped BMI of >30 category and by comparing this to the BMI of ≤25 category using a Mann-Whitney test. P < 0.05 was regarded as significant. Use of lipids GTL, other triglycerides, and fatty acids were sonicated into the media in a two-step fashion to ensure that the triglyceride stays in solution as described previously (53) . Briefly, pure GTL, GTO, LLP, LOP, and OOP were first directly sonicated into the PBS (concentration, 3 mM) using three 10-s pulses. This sonicate was promptly added to the medium (final concentration, 300 M). The only triglycerides requiring a solvent (DMSO) were those with two or more PA chains (PLP, GTP, and POP). These were prepared as follows: triglyceride stock (60 mM) was prepared in 100% DMSO by heating. These stocks were sonicated into PBS (concentration, 3 mM and 5% DMSO). These sonicates were quickly added to the medium (final concentration, 300 M and 0.5% DMSO). The experimental endpoints were not affected by this 0.1 to 0.5% DMSO in our system, as shown previously (6) . All animals were acclimatized for a minimum of 2 days before use. These were housed with a 12-hour light/dark cycle at room temperature, fed normal laboratory chow (ICR mice) or an experimental diet (see below), and were allowed to drink ad libitum. All experiments were approved by the Institutional Animal Care and Use Committee of the Mayo Clinic (Scottsdale, AZ). Ex vivo studies were done on pancreatic acini harvested from ICR mice (Charles River Laboratories, Wilmington, MA). In vivo studies were done on genetically obese male ob/ob (B6.V-lepob/J; the Jackson laboratory, Bar Harbor, ME) mice given an experimental diet below or ICR mice. The mice were 7 to 14 weeks of age at the time of studies. There were six to eight mice in each group. These were started immediately in ob/ob [B6.Cg-Lep ob /J; Jax Stock #00632, the Jackson laboratory (Bar Harbor, ME)] or C57bl/6J (to induce DIO) after weaning (age of 4 to 5 weeks), with an intent to change visceral fat composition. The diets are detailed in fig. S4 and were given ad libitum. Normal chow [Purina 5053 diet, LabDiet (www.labsupplytx.com/wp-content/uploads/2012/10/5053.pdf), which contains 5.0% of which ≈70% is UFA] and water were fed to control mice ad libitum. Mice were weighed periodically (1 to 4 weeks, depending on the rate of weight gain), and the body fat was measured using a Minispec Body Composition analyzer (Bruker Corporation, Billerica, MA) using nuclear magnetic resonance as previously described (27) . The models used were CER, and IL12 + IL-18 (IL12,18) induced pancreatitis in ob/ob mice, as detailed below. CER pancreatitis was also induced in C57bl/6 mice given the experimental diets or normal chow after 28 weeks. CER pancreatitis was induced by hourly intraperitoneal injections of CER (50 g/kg; Bachem AG, Bubendorf, Switzerland) × 12 doses on two consecutive days as described before (27) . IL12 (150 ng per dose per mouse; PeproTech) and IL18 (750 ng per dose per mouse; R&D Systems) pancreatitis was induced by giving IL12,18 as two doses 24 hours apart to lean ICR mice (10 to 14 weeks) as described previously (57) . Both agents were dissolved in PBS and given intraperitoneally. When studying the effect of intraperitoneal GTL (150 l) or GTP (150 mg) in these models, either agent was given 4 hours after the first CER injection or second IL12,18 injections. NEFA were administered into the peritoneal cavity of mice as described previously (7, 22, 28) . Carotid pulse distension (MouseOx Oximeter, STARR Life Sciences, Pittsburgh, PA) was used as a noninvasive measure of blood pressure. Mice were observed four times a day and were electively sacrificed at the end of the experiment using carbon dioxide, unless they became moribund prematurely. Terminal blood and tissue harvest was done if the animal was moribund and used in the analyses, unless specified for a different time point. Tissues were collected and preserved for mRNA (RNA later), activity (−80°C), histology and TUNEL staining (formalin), and Luminex and biochemical assays (serum) as described previously (6, 7, 28) . Histology and TUNEL staining Pancreas, kidney, and lung tissue were fixed with 10% neutral buffered formalin and processed for embedding in paraffin. Sections (5 m) were used for hematoxylin and eosin (H&E) staining and for TUNEL staining as described previously (5, 6, 27) . Digital images of section were captured with a digital microscope Axio Imager M2 or Axio Observer Z1 (Zeiss, Oberkochen, Germany) with a 20× objective with 10 random images per section and quantified as number of TUNEL-positive nuclei/high field of lung tissue. Pancreatic necrosis was quantified on whole-pancreas H&E-stained sections after being examined by a trained morphologist blinded to the sample. Briefly, all pancreatic parenchymal area was imaged using the PathScan Enabler IV slide scanner (Meyer Instruments, Houston, TX), and images were evaluated for total acinar area, total acinar necrosis, and acinar necrosis in direct proximity to the necrosed fat, i.e., peri-fat acinar necrosis as described previously (6, 27) in pixels for each pancreas. Percentage necrosis was reported as a percentage of the total acinar area for each pancreas. In vitro cell studies All data shown are from a three to five independent experiments. Triglycerides (e.g., GTL, LLP, LOP, and GTO) or NEFA were sonicated at 10× concentrations into the medium, and the only solvent used was 5% DMSO (as a 10× stock) for GTP and PLP or for palmitate, since these otherwise were insoluble and precipitated out in the medium. Solvents were otherwise avoided, since lipolysis in vivo takes place without solvents. These agents were added at a 1× concentration. Eight-to 12-week-old ICR mice (Charles River Laboratories, Wilmington, MA) were euthanized, and pancreatic acini were harvested as described (7, 28) . TEER and dextran permeability assay HUV-EC-C cells were grown as monolayers on a 0.4-m polyester Transwell permeable membrane (Corning, NY) precoated with type IV collagen and then used for trans-endothelial electrical resistance (TEER) measurement and permeability tracer flux assay. Before studies, cells were suspended in Hepes buffer (pH 7.4) (20 mM Hepes, 120 mM NaCl, 5 mM KCl, 1 mM MgCl 2 , 10 mM glucose, 10 mM sodium pyruvate, and 1 mM CaCl 2 ), and exposed to NEFA at the indicated concentration (60 M). TEER measurements were made using EVOM2 Epithelial Voltohmmeter with an STX2 electrode (World Precision Instruments, Sarasota, FL). Results were presented as absolute values (Ω·cm 2 ) after 2 hours of treatment. To evaluate the paracellular permeability of cultured HUV-EC-C cells, 10 kD of Dextran Alexa Fluor 647 (Thermo Fisher Scientific, Waltham, MA) was added to the upper chamber, and the increase in fluorescence was measured in the lower chamber 2 hours after stimulation with NEFA. For in vitro calcium ion flux and mitochondrial depolarization studies, cells were loaded with fura-2AM (5 g/ml; Molecular Probes, Invitrogen) and 5,5′,6,6′-tetrachloro-1,1′,3,3′-tetraethylbenzimidazolylcarbocyanine iodide (JC-1, 5 g/ml; Enzo Life Sciences, Farmingdale, NY) at 37°C for 30 min in the media that they were cultured or harvested in. After this, these were stored on ice till just before use. Just before the experiment, the cells were washed and suspended in Hepes buffer (pH 7.4) (20 mM Hepes, 120 mM NaCl, 5 mM KCl, 1 mM MgCl 2 , 10 mM glucose, and 10 mM sodium pyruvate) as previously described (6, 7, 28) . Triglycerides or NEFA were added after 2 min of stirring at a 1× concentration to the stirred cell suspensions in a quartz cuvette at 37°C, and changes in intracellular calcium concentrations were determined by alternate excitation at 340 and 380 nm, measuring emission at 510 nm. Mitochondrial inner membrane potential values (m) were determined by excitation at 490 nm and alternate measuring emission at 530 and 590 nm using the F2100 Hitachi Fluorescence Spectrophotometer. The data were collected every 10 s. The first 2 min of data were averaged and subtracted, and the net increase was plotted as a function of time or NEFA concentrations. In some studies, excess (i.e., 0.5%) fatty acid-free bovine serum albumin was added to neutralize and thus determine the changes attributable to the uNEFA. The assays (amylase, lipase, BUN, and calcium; Pointe Scientific, Canton, MI), DNA damage: Quant-iT PicoGreen dsDNA reagent (Life Technologies, Carlsbad, CA), LDH leakage using the Cytotoxicity Detection Kit (Roche, Mannheim, Germany), intracellular ATP measurement using ATP assay kit (CellTiter-Glo 2.0 Assay, Promega, Madison, USA), glycerol using free glycerol reagent, and standard (Sigma-Aldrich, St. Louis, MO) and colorimetric total free fatty acids using NEFA kit (FUJIFILM Wako Diagnostics, Mountain View, CA) were done as per the manufacturer's protocol (5, 6, 27) . uNEFA measurement by fluorescence uNEFAs were measured by the emission at 550/457 nm upon excitation at 375 nm and calculated as described in the ADIFAB2 kit (FFA Sciences LLC, CA, USA) protocol version 1.0 6-9-04. Standard concentration versus fluorescence ratio curves were made with known concentrations of either >99.5% purity LA, OA, or PA (1, 2, 5, 10, 20, 30, 50, and 100 M) in DMSO to have these all in the pure monomeric form. Then, the unknown (e.g., mouse serum) was measured and reported as uNEFA (micromolar concentrations), based on the fatty acid standard curve. Cytokine/chemokine assays Serum cytokine/chemokine protein levels were assayed with a MILLIPLEX MAP Mouse Adipokine Magnetic Bead Panel (Millipore, Burlington, MA) according to the manufacturer's recommendations on a Luminex 200 System (Life Technologies) and analyzed using the xPONENT software. These were done at the Hormone Assay and Analytical Services Core (Vanderbilt University Medical Center). For tissue triglyceride composition, the Folch method (76) was used, and the lipids recovered in the chloroform phase were separated by thin-layer chromatography using Silica Gel 60A plates. The triglyceride fraction was methylated using the BF3/methanol method (77) . These were analyzed using gas chromatography (Agilent 7890A) using a capillary column (SP2380, 0.25 mm by 30 m, 0.25-m film; Supelco, Bellefonte, PA) and helium as a carrier gas. Methyl esters of the fatty acids were identified by comparing their retention times to those of known standard flame ionization detection. Internal lipid standards for quantification included dipentadecanoyl phosphatidylcholine (C15:0), diheptadecanoin (C17:0), trieicosenoin (C20:1), and cholesteryl eicosenoate (C20:1). Serum-free fatty acids were extracted as described previously (78) in 30:70 heptane:isopropanol. This was acidified with 0.033 N H 2 SO 4 , and fatty acids were recovered in the heptane layer. Thin-layer chromatography and gas chromatography were done as described above. Dipentadecanoic acid (C15:0) was added as an internal standard. UFA amounts were calculated by adding individual C16:1, C18:1, C18:2, C18:3, C20:4, and C20:5 fatty acids. SFA amounts were calculated by adding individual C12:0, C14:0, C16:0, and C18:0 fatty acids. Western blotting was performed as described previously (54) . Briefly, cell lysate or fat pad lysate was prepared in radioimmunoprecipitation assay lysis buffer containing various protease inhibitors (complete and free of EDTA; Roche, Pleasanton, CA). Protein concentration (1 mg/ml) of the lysate was adjusted, boiled (10 min) in Laemmli buffer, and run on (10 mg per lane) 4 to 20% gradient polyacrylamide gels. Then, proteins were transferred from the gel to polyvinylidene difluoride membrane (Merck KGaA, Darmstadt, Germany) and blocked with 5% nonfat dry milk incubated with primary antibody of PNLIP (1:1000; Santa Cruz Biotechnology, Dallas, TX), IB- (1:1000; Santa Cruz Biotechnology), -tubulin (1:500; DSHB, University of Iowa, Iowa), cytochrome c (1:1000; Santa Cruz Biotechnology), and Cox-IV (1:1000; Thermo Fisher Scientific, Waltham, MA) separately. Then, blots were probed with horseradish peroxidaselabeled corresponding secondary antibodies (1:10,000; Millipore, Billerica, MA). Band intensity was visualized by chemiluminescence using the Pierce ECL 2 Western Blotting Substrate Kit (Thermo Fisher Scientific) and quantified by densitometry as described previously (79) . The hydrolysis of triglycerides by hPTL was carried out on a Nano ITC (TA Instruments-Waters LLC, New Castle, DE) instrument and through Nano ITCRun Software v3.5.6. The experimental setup was equipped with a computer-controlled micro-syringe injection device, which allowed to inject small amounts of stock solution of even 0.1 l to the sample chamber. Enzyme reaction was performed by two methods: either incremental injection of substrate to enzyme or continuous injection of enzyme to substrate. A total of 10 aliquots of freshly prepared, degassed GTL or LLP or LOP (0.4 mM, 5 l per injection) were injected from a rotating syringe of speed 350 rpm into the ITC sample chamber containing hPNLIP (12 nM, ~500 U/liter activity) equilibrated at 37°C. The interval between two injections was 300 s. Similarly, individual triglyceride into PBS or PBS into hPNLIP was run as a control. To confirm the specificity, LLL was injected into hPNLIP having orlistat (50 M orlistat; Cayman Chemical Company, Ann Arbor, MI) in a similar process. Degassed hPNLIP (45 l, 120 nM) was injected during 900 s from a rotating syringe into the ITC sample chamber containing LLL or LLP or LOP (0.1 mM) equilibrated at 37°C. The control experiment was performed by injecting PBS into hPNLIP. hPNLIP-mediated hydrolysis of LLL was confirmed using hPNLIP with orlistat in the same experimental setup. The ITC raw data were analyzed using the NanoAnalyze v3.10.0 software. Maximum velocity at saturating substrate concentration (V max ) was determined. Michaelis constant (K m ), the measure of dissociation/affinity of enzyme-substrate complex, was calculated from the substrate concentration at V max /2. Enzyme turnover rate (K cat ) was calculated by V max /enzyme concentration. Substrate to product formation rate (V) was calculated by the equation [K cat × enzyme concentration × substrate concentration]/[K m + substrate concentration]. For each experiment, we calculated V max , K m , K cat , and V values and summarized as means ± SEM from three independent experiments. From the raw data of continuous injection, enthalpy of the reaction was calculated from the delta heat change per mole. Interaction of triglyceride and hPNLIP followed by hydrolysis of triglyceride was graphically presented from three independent experiments. We followed the method described previously (54) . A total of 20 aliquots of degassed LA (0.7 mM, 2.5 l per injection from PBS, pH 7.4), OA (0.34 mM, 2.0 l per injection from PBS, pH 7.4), and PA (0.34 mM in 0.34% DMSO, 2.0 l per injection from PBS, pH 7.4) were injected (10 M per injection for LA and 4 M per injection for both OA and PA) from a rotating syringe (350 rpm) into the ITC sample chamber containing PBS equilibrated at 37°C. Double-distilled water was used in a reference electrode chamber. The interval between two injections was kept at 200 s. The control experiment was performed by injecting PBS into PBS and DMSO (0.34%) into PBS. Ultracentrifugation studies A total of 500 M C18:1, C18:2, and C18:3 were probe-sonicated into PBS (pH 7.4) at room temperature. For C16:0 and C18:0, 150 mM stocks were prepared in pure DMSO, and these were then serially sonicated and diluted down to 500 M (0.34% DMSO) in PBS (pH 7.4) at room temperature. A total of 4.8 ml of sample was loaded in OptiSeal 4.9-ml tubes (Beckman Instruments Inc., Palo Alto, CA) in a NVT90 fixed-angle rotor (Beckman Instruments Inc., Palo Alto, CA) and spun at 105g for 1 hour at room temperature in a Beckman Coulter Optima L-100 XP Ultracentrifuge (Beckman Coulter Inc., IN, USA). The bottom of the tubes was then pierced with a 25-gauge needle, and 1.0 ml of the liquid was removed. NEFA concentrations in these were measured using the colorimetric NEFA kit (FUJIFILM Wako Diagnostics, Mountain View, CA). Protein Data Bank structure 1LPA (interfacial activation of the lipase-procolipase complex by mixed micelles revealed by x-ray crystallography) was downloaded from the RCSB protein data bank (80) NY, 2019) was performed by selecting the four triglycerides as the ligands to be docked and designating the 1,2-didodecanoyl-sn-glycero-3-phosphocholine ligand from the 1LPA crystal structure as the centroid of the workspace ligand around which the receptor grid was generated. The ligand diameter midpoint box was constructed with dimensions of X, 10 Å; Y, 10 Å; and Z, 10 Å; while the receptor grid was generated with dimensions of X, 30 Å; Y, 30 Å; and Z, 30 Å. No constraints were imposed to ensure that the docking simulation was unbiased. The induced fit docking used the OPLS3e force field (82) . GlideScore docking energies were recorded in kilocalories per mole. Statistical analysis and graphical representation Independent variables for in vivo studies are shown in box plots in which the mean (dotted line), median (solid line), and in vitro data are shown as bar graphs reported as means ± SEM. Line graphs were used for continuous variables. All values are reported as means ± SD. Significance levels were evaluated at P < 0.05. Data for multiple groups were compared by one-way analysis of variance (ANOVA) versus controls, and values significantly different from controls were shown as (*) or with the P value mentioned above the corresponding conditions. When comparing two groups, a t test or Mann-Whitney test was used, depending on the normality of distribution and shown as (*) when significantly different; alternately, P values are shown in the graph. Graphing was done using SigmaPlot 12.5 (Systat Software Inc., San Jose, CA). Obesity: An important prognostic factor in acute pancreatitis Clinical predictors of severe acute pancreatitis: Value-adding the view from the end of the bed Abdominal and total adiposity and the risk of acute pancreatitis: A population-based prospective cohort study Free fatty acids in serum of patients with acute necrotizing or edematous pancreatitis Peripancreatic fat necrosis worsens acute pancreatitis independent of pancreatic necrosis via unsaturated fatty acids increased in human pancreatic necrosis collections Lipotoxicity causes multisystem organ failure and exacerbates acute pancreatitis in obesity Pancreatic triglyceride lipase mediates lipotoxic systemic inflammation Implementation of the Asia-Pacific guidelines of obesity classification on the APACHE-O scoring system and its role in the prediction of outcomes of acute pancreatitis: A study from India Influence of obesity on the severity and clinical outcome of acute pancreatitis The clinical relevance of obesity in acute pancreatitis: Targeted systematic reviews Potential role of adipose tissue and its hormones in burns and critically iII patients ADHERE Scientific Advisory Committee and Investigators, An obesity paradox in acute heart failure: Analysis of body mass index and inhospital mortality for 108,927 patients in the Acute Decompensated Heart Failure National Registry Body mass index predicts perioperative complications following orthopaedic trauma surgery: An ACS-NSQIP analysis The obesity paradox is still there: A risk analysis of over 15 000 cardiosurgical patients based on body mass index Is body mass index associated with outcomes of mechanically ventilated adult patients in intensive critical units? A systematic review and meta-analysis Association between the pancreatic enzyme level and organ failure in trauma patients Clinical significance of increased lipase levels on admission to the ICU Hyperamylasemia after cardiac surgery. Incidence, significance, and management Simultaneous measurement of the rates of appearance of palmitic and linoleic acid in critically ill infants Leukotoxin, a linoleate epoxide: Its implication in the late death of patients with extensive burns ACE2 expression in pancreas may cause pancreatic damage after SARS-CoV-2 infection Lipotoxicity in COVID-19 Study Group, Mortality from coronavirus disease 2019 increases with unsaturated fat and may be reduced by early calcium and albumin supplementation COVID-19 infection alters kynurenine and fatty acid metabolism, correlating with IL-6 levels and renal status Lipotoxicity and cytokine storm in severe acute pancreatitis and COVID-19 Relationships between composition of major fatty acids and fat distribution and insulin resistance in Japanese Increase in adipose tissue linoleic acid of US adults in the last half century Lipolysis of visceral adipocyte triglyceride by pancreatic lipases converts mild acute pancreatitis to severe pancreatitis independent of necrosis and inflammation Ringer's lactate prevents early organ failure by providing extracellular calcium Fatty acid composition of adipose tissue and blood in humans and its use as a biomarker of dietary intake Decreased severity in recurrent versus initial episodes of acute pancreatitis Early inflammatory response in acute pancreatitis is little affected by body mass index Does obesity confer an increased risk and/or more severe course of post-ERCP pancreatitis?: A retrospective, multicenter study Does the presence of obesity and/or metabolic syndrome affect the course of acute pancreatitis?: A prospective study The panc 3 score: A rapid and accurate test for predicting severity on presentation in acute pancreatitis Obesity as a predictor of severity in acute pancreatitis Is obesity a significant prognostic factor in acute pancreatitis? Obesity: A prognostic factor of severity in acute pancreatitis Combination of APACHE-II score and an obesity score (APACHE-O) for the prediction of severe acute pancreatitis Obesity increases the severity of acute pancreatitis: Performance of APACHE-O score and correlation with the inflammatory response Impact of body overweight and class I, II and III obesity on the outcome of acute biliary pancreatitis Obesity and fat distribution imply a greater systemic inflammatory response and a worse prognosis in acute pancreatitis Early jejunal feeding initiation and clinical outcomes in patients with severe acute pancreatitis Effect of obesity and decompressive laparotomy on mortality in acute pancreatitis requiring intensive care unit admission Obesity as a risk factor for severe acute pancreatitis patients Bisap-O: Obesity included in score BISAP to improve prediction of severity in acute pancreatitis Obesity: A risk factor for severe acute biliary and alcoholic pancreatitis APACHE system is better than Ranson system in the prediction of severity of acute pancreatitis A large volume of visceral adipose tissue leads to severe acute pancreatitis Prevention of severe acute pancreatitis with octreotide in obese patients: A prospective multi-center randomized controlled trial Epidemiology and outcome of acute pancreatitis The epidemiology of pancreatitis and pancreatic cancer Changes in consumption of omega-3 and omega-6 fatty acids in the United States during the 20th century Carboxyl ester lipase may not mediate lipotoxic injury during severe acute pancreatitis Multimodal transgastric local pancreatic hypothermia reduces severity of acute pancreatitis in rats and increases survival Prognostic factors in acute pancreatitis A single-centre double-blind trial of trasylol therapy in primary acute pancreatitis Effect of diet-induced obesity on acute pancreatitis induced by administration of interleukin-12 plus interleukin-18 in mice Action of secretagogues on a new preparation of functionally intact, isolated pancreatic acini Stereoselectivity of lipases. II. Stereoselective hydrolysis of triglycerides by gastric and pancreatic lipases Inhibition of pancreatic lipase and triacylglycerol intestinal absorption by a Pinhão Coat (Araucaria angustifolia) extract rich in condensed tannin Interfacial activation of the lipase-procolipase complex by mixed micelles revealed by x-ray crystallography The 2.46 A resolution structure of the pancreatic lipase-colipase complex inhibited by a C11 alkyl phosphonate High-throughput virtual screening with e-pharmacophore and molecular simulations study in the designing of pancreatic lipase inhibitors The relationship of waist circumference and BMI to visceral, subcutaneous, and total body fat: Sex and race differences The inhibition of the GTPase activating protein-Ha-ras interaction by acidic lipids is due to physical association of the C-terminal domain of the GTPase activating protein with micellar structures The interplay between inflammation, coagulation and endothelial injury in the early phase of acute pancreatitis: Clinical implications WHO Expert Consultation, Appropriate body-mass index for Asian populations and its implications for policy and intervention strategies The paradox of low body mass index and high body fat percentage among Chinese, Malays and Indians in Singapore Acute pancreatitis: Value of CT in establishing prognosis Body size, body composition and fat distribution: Comparative analysis of European, Maori, Pacific Island and Asian Indian adults Plank, BMI, fat and muscle differences in urban women of five ethnicities from two countries Dutch Pancreatitis Study Group, The association of computed tomography-assessed body composition with mortality in patients with necrotizing pancreatitis Role for stearoyl-CoA desaturase-1 in leptin-mediated weight loss Crystallization, polymorphism, and binary phase behavior of model enantiopure and racemic triacylglycerols The activation of porcine pancreatic lipase by cis-unsaturated fatty acids A simple method for the isolation and purification of total lipides from animal tissues Preparation of fatty acid methyl esters and dimethylacetals from lipids with boron fluoride-methanol A gas-liquid chromatographic assay for plasma free fatty acids Transcriptional regulation of CXC-ELR chemokines KC and MIP-2 in mouse pancreatic acini The Protein Data Bank PubChem 2019 update: Improved access to chemical data OPLS3: A force field providing broad coverage of drug-like small molecules and proteins Funding was obtained by V.P.S. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data Adipose saturation reduces lipotoxic systemic inflammation and explains the obesity paradox Acknowledgments: We thank J. N. Singh and R. Cannon for working with us on the illustration shown in Fig. 7 . Funding: This project was supported by R01DK092460 and R01DK119646 from the NIDDK and PR151612 and PR191945 from the DOD (to V.P.S.). Author contributions: V.P.S. designed, supervised, and conceptualized the study. Acquisition and Supplementary material for this article is available at http://advances.sciencemag.org/cgi/ content/full/7/5/eabd6449/DC1 View/request a protocol for this paper from Bio-protocol.