4000-520-616
欢迎来到免疫在线!(蚂蚁淘生物旗下平台)  请登录 |  免费注册 |  询价篮
主营:靶向毒素
咨询热线电话
4000-520-616
当前位置: 首页 > 新闻动态 >
新闻详情
...of a New Generation of Anti-inflammatory Drugs Targeting...
来自 : 发布时间:2024-05-02
DREAM-in-CDM Approach and Identification of a New Generation of Anti-inflammatory Drugs Targeting mPGES-1 AbstractMicrosomal prostaglandin E2 synthase-1 (mPGES-1) is known as an ideal target for next generation of anti-inflammatory drugs without the side effects of currently available anti-inflammatory drugs. However, there has been no clinically promising mPGES-1 inhibitor identified through traditional drug discovery and development route. Here we report a new approach, called DREAM-in-CDM (Drug Repurposing Effort Applying Integrated Modeling-in vitro/vivo-Clinical Data Mining), to identify an FDA-approved drug suitable for use as an effective analgesic targeting mPGES-1. The DREAM-in-CDM approach consists of three steps: computational screening of FDA-approved drugs; in vitro and/or in vivo assays; and clinical data mining. By using the DREAM-in-CDM approach, lapatinib has been identified as a promising mPGES-1 inhibitor which may have significant anti-inflammatory effects to relieve various forms of pain and possibly treat various inflammation conditions involved in other inflammation-related diseases such as the lung inflammation caused by the newly identified COVID-19. We anticipate that the DREAM-in-CDM approach will be used to repurpose FDA-approved drugs for various new therapeutic indications associated with new targets. IntroductionA variety of human diseases, such as the newly identified coronavirus disease 2019 (COVID-19), various forms of pain, cardiovascular diseases, neurodegenerative diseases, and cancers1,2,3, involve serious inflammation conditions without a truly effective and safe anti-inflammatory drug to suppress. Prostaglandin E2 (PGE2) serves as the principal proinflammatory agent involved in various inflammation-related diseases4. The biosynthesis5 of PGE2 starts from a polyunsaturated fatty acid, known as arachidonic acid (AA), freed from phospholipids. The AA is first converted to prostaglandin H2 (PGH2) by cyclooxygenase (COX)-1 and COX-25. Then, PGH2 is converted to PGE2 by prostaglandin E synthase (PGES)6. The traditional nonsteroidal anti-inflammatory drugs (NSAIDs) either weakly and non-selectively inhibit both COX-1 and COX-2 or potently and selectively inhibit COX-2. All these traditional drugs have a number of serious adverse side effects, including the increased risk of fatal heart attack or stroke and stomach or intestinal bleeding etc., because the COX-1/2 inhibition also blocks the synthesis of all of the other prostaglandins (PGs), including PGI2, PGD2, PGF2伪, and TXA2 that are required physiologically, synthesized from PGH27. In comparison with the COX-1/2 inhibition, inhibition of the downstream enzyme, known as microsomal PGES-1 (mPGES-1), will only block the PGE2 overproduction without blocking the biosynthesis of the other physiologically required PGs as confirmed by reported knock-out studies8,9. So, mPGEA-1, as an inducible enzyme under the inflammation conditions, is a much more promising target for a next generation of anti-inflammatory drugs. One may reasonably anticipate that a highly selective mPGES-1 inhibitor will fully retain the desired anti-inflammatory effects of the COX-1/2 inhibition, but lack of the unwanted adverse effects produced by the inhibition of COX-1/2.In fact, various inhibitors of human mPGES-1 have been reported in the literature for development of the next-generation inflammatory drugs10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31. Unfortunately, it is still rare to see that a potent inhibitor of human mPGES-1 can also potently inhibit mouse or rat mPGES-1. This problem prevents using well-established mouse/rat models of inflammation, pain, and other diseases for preclinical drug development studies32. It would be extremely challenging for the mPGES-1-based drug development to follow the traditional drug discovery and development route (which starts from lead identification and optimization based on various in vitro and in vivo assays in terms of the activity and toxicity etc.). Here we report a new approach, called DREAM-in-CDM (Drug Repurposing Effort Applying Integrated Modeling-in vitro/vivo-Clinical Data Mining), to identify an FDA-approved drug suitable for use as an effective analgesic targeting mPGES-1. The DREAM-in-CDM approach consists of three steps: (1) computational modeling to predict which FDA-approved drugs may favorably bind with the desirable drug target (mPGES-1 in this project); (2) in vitro and/or in vivo assays to validate the computational predictions; (3) clinical data mining to confirm the efficacy associated with the required clinical end points for the new therapeutic indication. By using the DREAM-in-CDM approach, multiple FDA-approved drugs (that can be used orally), including lapatinib, have been identified as truly promising mPGES-1 inhibitors that may be repurposed to treat various inflammation-related diseases.ResultsPredictions from computational modelingOn the basis of our previous computational modeling studies32,33,34, we recently obtained the open conformation of human mPGES-1 through fully relaxed molecular dynamics simulations35. The modeled open conformation of human mPGES-1 was used further to virtually screen FDA-approved drugs for their potential binding with mPGES-1 in this study. The virtual screening was performed by using a multiple-step computational screening protocol (see the Materials and Methods section below). Based on the virtual screening (including molecular docking, energy minimization, MM-PBSA binding free energy calculation, and manual checking of the binding poses for the top-50 compounds), 15 FDA-approved drugs (that can be used orally), including lapatinib, were predicted to be inhibitors of mPGES-1. For further computational testing, we also tested the binding of the 15 drugs with the high-resolution crystal structure (4AL0 with a high resolution at 1.2鈥壝? of mPGES-1 in which co-factor glutathione (GSH) was removed from the binding site before molecular docking, confirming that these drugs can bind to the GSH-binding site of the crystal structure (see Fig.聽1A for the binding mode of lapatinib). So, both the computationally modeled structure and the X-ray crystal structure of mPGES-1 can reasonably accommodate these drugs in the same GSH-binding site.Figure 1(A) Modelled binding structure of lapatinib (cyan sticks) with mPGES-1 (yellow cartoon) obtained by using the crystal structure (4AL0). The cap domain is colored in green and some key residues interacting with lapatinib are showed in sticks. Hydrogen bond is showed in black dashed line. (B) Dose-dependent inhibition of lapatinib against mPGES-1. (C) Inhibition of mPGES-1 with lapatinib at constant 1鈥壜礛 concentration and GSH at different concentrations, showing that lapatinib is a GSH-competitive inhibitor.Full size imageIn addition, we also examined whether lapatinib can fit the active site cavity of COX-2 by performing molecular docking in comparison with celecoxib (a well-known potent inhibitor of COX-2 in clinical use). It turned out that the binding pocket of celecoxib was too tight for lapatinib to fit into. Based on the molecular docking, lapatinib was not expected to be an effective inhibitor of COX-2.In vitro activityThe above computational predictions were followed by in vitro activity assays for their actual inhibitory activity against human mPGES-1 using an ELISA assay described in our previous reports32,33,34. All compounds for the FDA-approved drugs from the Enzo Compound Library were provided with a purity of 95% or better36. Specifically, 15 FDA-approved drugs (see Fig.聽S1 of the Supporting Information for the molecular structures) were assayed for their inhibitory activity against human mPGES-1. According to the initial single-concentration (at 10鈥壜礛) screening, the 15 drugs inhibited the mPGES-1 activity by 4% to 99%, and 7 out of the 15 drugs at 10鈥壜礛 inhibited the mPGES-1 activity by 44% or more. The top-7 drugs were assayed further for their IC50 values (see Table聽1), with lapatinib being the most potent one (IC50鈥?鈥?.8鈥壜礛 or 800鈥塶M). In addition, some other FDA-approved drugs can also significantly inhibit human mPGES-1, but with relatively higher IC50 values (lower potency). So, we first focused on lapatinib in further experimental tests in this investigation.Table 1 The FDA-approved drugs selected from virtual screening and tested for their in vitro inhibitory activity against human mPGES-1.Full size tableFurther, we determined in vitro inhibitory activity of lapatinib against mouse mPGES-1 and obtained that IC50鈥?鈥?2鈥壜礛. So, lapatinib has a 15-fold lower inhibitory activity against mouse mPGES-1 compared to human mPGES-1 (IC50鈥?鈥?.8鈥壜礛).In addition, lapatinib was assayed for its potential inhibitory activity against mixed COX-1 and COX-2 (denoted as COX-1/2) with equal amounts of COX-1 and COX-2 in terms of the enzyme activities. Notably, lapatinib at a concentration of 100鈥壜礛 did not significantly inhibit COX-1/2, which confirmed the above computational prediction.Binding mode analysisDepicted in Fig.聽1 are the computationally modeled protein-lapatinib binding structure (Fig.聽1A, showing a favorable hydrogen bond between oxygen on the furan ring of lapatinib and the hydroxyl group of S127 side chain) obtained by using the crystal structure (4AL0) and in vitro activity data (Fig.聽1B) obtained for lapatinib. To examine whether lapatinib competes with GSH in binding mPGES-1, we tested the mPGES-1 enzyme activity (enzymatic reaction velocity V) with various GSH concentrations in the presence and absence of 1鈥壜礛 lapatinib to determine the Lineweaver-Burk plots (Fig.聽1C). As well-known, by using a Lineweaver-Burk plot, one can obtain the Michaelis-Menten constant (KM) from the x-intercept (i.e. the x-intercept鈥?鈥夆垝1/KM) and Vmax from the y-intercept (i.e. the y-intercept = 1/Vmax). Thus, we obtained KM鈥?鈥?.1鈥塵M (from the x-intercept of 鈭?.4700 in Fig.聽1C) in the presence of 1鈥壜礛 lapatinib and KM鈥?鈥?.72鈥塵M (from the x-intercept of 鈭?.393 in Fig.聽1C) in the absence of lapatinib. So, 1鈥壜礛 lapatinib increased KM of human mPGES-1 associated with GSH from 0.72鈥塵M to 2.1鈥塵M. As seen in Fig.聽1C, the y-intercept = 0.15鈥壜扁€?.02 in the absence of lapatinib and the y-intercept = 0.18鈥壜扁€?.02 in the presence of 1鈥壜礛 lapatinib; there was no significant difference in the y-intercept within the experimental errors. Hence, we can conclude that 1鈥壜礛 lapatinib did not significantly change the Vmax, but significantly increased the KM with respect to GSH, which is the well-known feature of the competitive inhibition. So, the experimental data depicted in Fig.聽1C support that lapatinib indeed competes with co-factor GSH, which is consistent with the computationally modeled binding mode in which lapatinib occupies the GSH-binding site.In vivo anti-inflammatory activity of lapatinib in miceTo examine the anti-inflammatory potential of lapatinib, we determined the in vivo effectiveness of lapatinib using the most popularly used mouse air-pouch model of inflammation in comparison with celecoxib (a positive control). According to the in vivo data depicted in Fig.聽2, lapatinib significantly and dose-dependently decreased the air-pouch PGE2 level in carrageenan-treated mice. In comparison with the positive control (celecoxib) at a dose of 50鈥塵g/kg (PO), lapatinib at 50鈥塵g/kg (PO) is less effective, as expected. Interestingly, there was no significant difference in the in vivo potency between 100鈥塵g/kg lapatinib (PO) and 50鈥塵g/kg celecoxib (PO), as shown in Fig.聽2.Figure 2Data from in vivo assays using the mouse air-pouch model (n鈥?鈥? for each group) with lapatinib or celecoxib given PO. The air-pouch fluid collected from mice were analyzed by ELISA assay for the PGE2 concentrations. GraphPad Prism 7 software (GraphPad Software, La Jolla, CA) was used to perform the one-way analysis of variance (ANOVA) with post hoc tests.Full size imageOne can estimate the human equivalent dose (HED) associated with the highly effective mouse dose of 100鈥塵g/kg by using the commonly used practice guide37 for dose conversion between different species:$${rm{HED}}=({{rm{D}}}_{{rm{animal}}}/{{rm{K}}}_{{rm{animal}}})={{rm{D}}}_{{rm{mouse}}}/12.3$$ in which Kanimal is the correction factor (which is 12.3 for mouse) when the compound is equally potent against the human and animal target proteins. Further in consideration of the species difference in the in vitro potency, Eq. (1) can be extended in this study as the following:$${rm{HED}}=({{rm{D}}}_{{rm{mouse}}}/12.3)times {{rm{IC}}}_{50}({rm{human}},{rm{mPGES}}-1)/{{rm{IC}}}_{50}({rm{mouse}},{rm{mPGES}}-1)$$ According to Eq. (2), one can estimate the HED associated with the mouse dose of 100鈥塵g/kg: HED鈥?鈥?100/12.3) 脳 (0.8/12) = ~0.54鈥塵g/kg (or about 35鈥塵g per dose for human with an average body weight of 65鈥塳g). The in vivo data and the effective HED estimate suggest that lapatinib may serve as an effective anti-inflammatory drug to treat inflammation-related diseases for humans. It should be noted that this is merely an estimation based on the commonly used practice. The actual minimum effective dose for human can only be determined through actual clinical trials for further drug repurposing studies in the future.Insights from clinical data miningAs an FDA-approved drug, lapatinib was tested in many clinical trials. In clinical data mining, we would like to know whether lapatinib can reach the effective in vivo concentration in human body comparable to the IC50 value (~800鈥塶M). To address this question, we collected the pharmacokinetic (PK) data, particularly the maximum drug concentration (Cmax) values, of lapatinib from five different clinical trials and estimated the corresponding Cmax values in different tissues/organs. According to the data collected in Table聽2, the plasma Cmax of lapatinib in its usual dosage forms should be well above its IC50 against human mPGES-1. In addition, the reported ratios of Cmax values in various organs of rat to Cmax in rat plasma, along with the Cmax values in human plasma, were used to predict the Cmax values in the corresponding organs of human corresponding to the Cmax values in human plasma. The predicted Cmax values in various human organs are also listed in Table聽2. As seen in Table聽2, lapatinib may have much higher concentrations in various human organs compared to the corresponding concentration in human plasma. In particular, Cmax is predicted to be as high as ~2鈥壜礛 ( IC50) in lung and ~1.3鈥壜礛 ( IC50) in kidney at a dose of only 50鈥塵g. Notably, kidney is rich of mPGES-132. All of these PK data suggest that lapatinib can serve as an effective anti-inflammatory drug to treat inflammation and pain as well as many other inflammation-related diseases.Table 2 Cmax (碌M) of lapatinib in human plasma from clinical pharmacokinetic data and the predicted Cmax (碌M) in various human organs associated with various doses.Full size tableDiscussionAs well known, mPGES-1 is very promising, but extremely difficult, target for practical drug development. A major difficulty exists in the difference in the PGH2-binding site of the enzyme between different species. For this reason, majority of the previously reported human mPGES-1 inhibitors are inactive against mouse or rat mPGES-1 such that one cannot use the well-established mouse/rat models of inflammation, pain, and other diseases for preclinical studies. Hence, it is interesting to explore novel inhibitors that can bind to the GSH-binding site which is more conserved for different species. This is the first report of an mPGES-1 inhibitor competitive with GSH. The competitive inhibition has been supported by the in vitro experiment. Further effort in this topic will be needed to obtain a co-crystal structure of mPGES-1 with lapatinib.Our structure-based virtual screening of FDA-approved drugs, followed by in vitro and in vivo activity assays, has demonstrated that multiple FDA-approved drugs, including lapatinib, can significantly inhibit mPGES-1. Thus, these FDA-approved drugs could also have anti-inflammatory and analgesic effects, depending on their actual pharmacokinetic profiles associated with the actual dosage forms. Further clinical data mining revealed that lapatinib has a favorable pharmacokinetic profile to ensure that the lapatinib concentrations in various tissues can reach the levels that are significantly higher than the IC50 value after orally taking a lower dose of lapatinib. So, lapatinib may serve as a highly desired anti-inflammatory drug targeting mPGES-1 to treat inflammation and pain as well as a variety of other inflammation-related diseases. For example, the recently identified COVID-19 causes serious lung inflammation. It would be interesting to test whether lapatinib can be used to effectively suppress the inflammation caused by COVID-19 while an antiviral drug (such as remdesivir or chloroquine or hydroxylchloroquine or any other antiviral drug to be approved by the FDA) is used to kill the virus in the body.It has been known that lapatinib has its anti-cancer activity through inhibition of epidermal growth factor receptor (EGFR) and human epidermal growth factor receptor 2 (HER2). So far, lapatinib has been used only for its anti-cancer activity in clinical practice. Hence, further clinical trials will be required before FDA approves repurposing lapatinib as an anti-inflammatory drug to treat inflammation and pain as well as a variety of inflammation-related diseases. Notably, reported phase I clinical trials in non-cancer patients38 indicated that lapatinib did not cause any serious adverse effects or significantly increase other common adverse effects, suggesting that lapatinib has the desirable safety as a novel anti-inflammatory drug for both cancer and non-cancer patients. In another Phase I dose-escalation study39, the highest dose of lapatinib used was as high as 7000鈥塵g per day in twice-daily dosing with no dose-limiting toxicity (DLT) found, further demonstrating the safety of lapatinib. On the other hand, it is possible that lapatinib has side-effects that are not relevant in patients with cancer, but that can be important in patients with inflammatory diseases. The risk/benefit of drugs might be dependent on the clinical condition that is taken into consideration. So, it would be interesting to conduct drug repurposing clinical trials on lapatinib for its analgesic effects, because only clinical trials can determine its efficacy, as well as the safety, as an anti-inflammatory drug.Concerning whether the mPGES-1 inhibition will indirectly increase the production of other PGs while decreasing the PGE2 production, there have been a variety of mPGES-1 knockout studies in mice. The genetic inactivation of mPGES-1 has been shown to result in the shunting of the PGH2 substrate to other terminal PG synthases, thereby increasing the synthesis of other PGs. Different reports showed that in explants of peritoneal macrophages isolated from mPGES-1-deficient mice that the reduction in PGE2 synthesis was accompanied by an increase in PGI2 and TXA2 stable metabolites40, increased PGI2 synthesis without increased TXA2 synthesis41, or no change in PGI2 and TXA2 synthesis42. Although PGD2 and PGF2伪 production from peritoneal macrophages have also been shown to be altered following mPGES-1 deficiency, changes in these products have been shown to be less significant than the effects on PGI2 and TXA240. Although a variety of studies using in vitro cell culture or cell explant culture systems have shown the potential for the shunting of PGH2 substrate following mPGES-1 inactivation, the causative effects of substrate shunting in vivo is less well defined. An adverse effect on cardiovascular function resulting from increases in thrombogenic TXA2 production or decreases in cardioprotective PGI2 is one area where changes in the PG synthetic profile following mPGES-1 inactivation has been of particular concern. However, an in vivo study examining cardiovascular function showed no increase in TXA2 synthesis, thrombosis, or blood pressure in mPGES-1-deficient mice, and an increase in cardioprotective PGI2 production, thereby suggesting the potential for mPGES-1 inhibitors to be effective anti-inflammatory agents without adverse cardiovascular effects7. There was also concern that reduced synthesis of PGE2 due to inactivation of mPGES-1 may also produce adverse effects in tissues where PGE2 is responsible for maintaining physiological function. Although the genetic deficiency of mPGES-1 was reported to decrease gastric PGE2 levels by approximately 80%42, the deficiency of mPGES-1 is not sufficient to induce gastric lesions. Additionally, our previous study in wild-type mice using a recently designed highly selective mPGES-1 inhibitor (4b)32 suggested that mPGES-1 should be a much safer target than COX-2, because 4b at a high dose (up to 5,000鈥塵g/kg) administered (PO) did not cause any toxicity in mice while 50鈥塵g/kg celecoxib (a selective COX-2 inhibitor) administered PO had significant toxicity for stomach and other issues of mice.Further, the scaffold of the lapatinib structure may be used as a new starting point for future rational design of a lapatinib analog which can more potently inhibit mPGES-1 without inhibiting EGFR/HER2. Such a lapatinib analog may serve as an improved analgesic compared to lapatinib itself in terms of both the efficacy and side effects. Similarly, the scaffolds of other FDA-approved drugs identified as mPGES-1 inhibitors (with the IC50 values listed in Table聽1) may also be used as new starting points for rational design of their analogs that can more potently and selectively inhibit mPGES-1.In addition, the general DREAM-in-CDM approach may also be useful for repurposing FDA-approved drugs for other new therapeutic indications associated with new targets.Materials and MethodsIn silico simulationThe virtual screening of the FDA-approved drug from the Enzo Compound Library (http://www.enzolifesciences.com/BML-2843/screen-well-fda-approved-drug-library-v2/) in this study was based on the use of the recently reported open-conformation35 of human mPGES-1 and a revised version (with a unique feature) of our previously described computational protocol33. Prior to the virtual screening, the 3D structures of the FDA-approved drugs were generated by using OpenBabel43. Starting from the prepared structures, the computational protocol includes the use of AutoDock Vina 1.1.2 software44 for rigid molecular docking. A 15鈥壝?脳 15鈥壝?脳 15鈥壝?docking box containing the original GSH-binding site of mPGES-1 was chosen as the targeted binding site and all other parameters were set to default. Only the top-ranked pose of the ligand was further evaluated using the AMBER16 software45. In all the subsequent simulations using the AMBER16, only ligand atoms and hydrogen atoms of the protein within 4鈥壝?of the ligand were set flexible. For all the binding complexes, a four-step procedure, including 4,000 steps of energy minimization (1,000 steps with the steepest descend method and then 3,000 steps with the conjugate gradient method), 20鈥塸s of MD simulation (temperature set to 300鈥塊, time step was set to 2鈥塮s, bonds involving hydrogen are constrained using SHAKE algorithm46 6,000 steps of energy-minimization (2,000 steps with the steepest descend method and then 4,000 steps with the conjugate gradient method), and the molecular mechanics-Poisson-Boltzmann surface area (MM-PBSA) calculations47 (mbondi2 radii was used, parameter igb was set to 5, ionic strength was set to 0.1鈥塵M, and external dielectric constant was set to 78.5.) for estimating the binding free energies. This multiple-step computational procedure is similar to the known binding estimation after refinement (BEAR)48,49 protocol. However, for a major difference between our current computational protocol and the BEAR protocol48,49, the free energy difference between the lowest-energy conformation of the free ligand and the conformation of the ligand in the mPGES-1-ligand binding complex was also calculated and added to the total binding free energy. Technically, the MM-PBSA binding free energy (螖Gbind) was evaluated as$${{rm{DG}}}_{{rm{bind}}}={rm{DG}},({rm{complex}})-{rm{DG}},({rm{free}},{rm{ligand}})-{rm{DG}},({rm{protein}})$$ In Eq. (3), as usual, 螖G (complex) and 螖G (protein) are the free energies of the protein-ligand complex and the free protein. However, 螖G (free ligand) is the free energy of the ligand in its free ligand conformation, instead of the ligand conformation existing in the protein-ligand complex. In this way, the calculated binding free energies are expected to be more reasonable, because they appropriately account for the conformational free energy change from the free-ligand conformation to the conformation in the protein-ligand binding complex. The compounds were then ranked by the estimated binding free energies and screened by the PAINS filter. According to the calculated binding free energies, the top-50 compounds were manually checked to remove the ones with clearly unreasonable ligand conformations, leading to selection of top-15 compounds.In addition, the top-15 compounds were tested further for their binding with mPGES-1 in the high-resolution crystal structure (4AL0 with a high resolution at 1.2鈥壝? by using the same computational protocol (molecular docking, energy minimizations, MD simulations, and binding free energy calculations). Prior to the molecular docking, co-factor glutathione (GSH) in the crystal structure was removed from the binding site.Molecular docking was also carried out for possible binding of lapatinib with COX-2 by using a reported X-ray crystal structure (PDB 5JW1)50.In vitro activity assaysThe protocol for the protein preparation and in vitro activity assays were the same as described previously51,52,53,54. The enzyme activity assays were performed on ice in 1.5鈥塵l microfuge tubes by using the expressed human or mouse mPGES-1. The reaction mixture contained: 0.2鈥塎 Na2HPO4/NaH2PO4, pH 7.2, 10鈥壜礚; 0.1鈥塎 GSH, 2.5鈥壜礚; diluted microsomal enzyme (80鈥壜礸/mL), 1鈥壜礚; PGH2 (0.31鈥塵M in DMF), 5鈥壜礚; 1鈥壜礚 inhibitor; and H2O in a final reaction volume of 100鈥壜礚. PGH2 was stored in dry ice and used to initiate the reaction. Compounds were incubated with the enzyme for 15鈥塵in at room temperature before the addition of cold PGH2 (1鈥壜礛 final) to initiate the enzyme reaction. After 30鈥塻, 10鈥壜礚 of SnCl2 (40鈥塵g/mL) in ethanol was added to stop the reaction. The non-enzymatic conversion of PGH2 to PGE2 was performed in the same buffer devoid of enzyme. The reaction mixture was placed on ice until PGE2 production was determined by the PGE2 enzyme immunoassay as described earlier. IC50 values of the inhibitors were calculated by using the GraphPad Prism 7.The inhibitory activity of lapatinib (the most active mPGES-1 inhibitor identified in this study) against COX isoenzymes was determined by using the COX (ovine/human) inhibitor screening assay kit purchased from Cayman Chemical Company (Ann Arbor, MI). We used a mixture of equal amounts of purified COX-1 and COX-2 proteins and followed the protocol recommended by the vendor.In vivo activity testsFor in vivo tests, additional lapatinib sample (10鈥塯 at a purity of 99% determined by HPLC) was ordered from the LC laboratories (Woburn, MA). All the animal experiments were conducted in our animal laboratories in the University of Kentucky鈥檚 Division of Laboratory Animal Resources (DLAR) facility (PHS assurance number A3336-01; USDA number 61-R-0002; AAALAC, Intl. Unit # 13) following the guidelines and regulations of the AAALAC and National Institutes of Health (NIH). The animal procedure used in this project was approved by the University of Kentucky鈥檚 Institutional Animal Care and Use Committee (IACUC). The air-pouch model of inflammation23,55 is widely used for determining the in vivo effectiveness of inhibitors of prostaglandin synthesis. Air pouches were produced by duplicate injections of 3鈥塵L of sterile air under the skin on the back of mice. After the formation of the air-pouch, a single injection of the inflammatory agent carrageenan into the pouch resulted in the recruitment of inflammatory cells and the production of a fluid exudate containing significant levels of PGE2 (an inflammatory marker) produced primarily by activities of COX-2 and mPGES-1. Then, the mice were treated PO (oral gavage) with a single dose of lapatinib, celecoxib, or vehicle for 24鈥塰ours prior to collection of air-pouch fluid samples. The air-pouch fluid samples were analyzed for PGE2 by the same ELISA method used in the in vitro enzyme activity assay mentioned above.Statistical analysisGraphPad Prism 7 software (GraphPad Software, La Jolla, CA) was used to perform the one-way analysis of variance (ANOVA) with post hoc testing, allowing us to examine the significance of the difference in the in vivo activity data between each pair of dose conditions. p鈥?lt;鈥?.05 was considered statistically significant. References1.Radmark, O. Samuelsson, B. Microsomal prostaglandin E synthase-1 and 5-lipoxygenase: potential drug targets in cancer. J. Intern. Med. 268, 5鈥?4 (2010).CAS聽 PubMed聽Google Scholar聽 2.Hanaka, H. et al. Microsomal prostaglandin E synthase 1 determines tumor growth in vivo of prostate and lung cancer cells. Proc. Natl. Acad. Sci. USA 106, 18757鈥?8762 (2009).ADS聽 CAS聽 Article聽Google Scholar聽 3.Koeberle, A. Werz, O. Perspective of microsomal prostaglandin E2 synthase-1 as drug target in inflammation-related disorders. Biochem. Pharmacol. 98, 1鈥?5 (2015).CAS聽 Article聽Google Scholar聽 4.Serhan, C. N. Levy, B. Success of prostaglandin E-2 in structure-function is a challenge for structure-based therapeutics. Proc. Natl. Acad. Sci. USA 100, 8609鈥?611, https://doi.org/10.1073/pnas.1733589100 (2003).ADS聽 CAS聽 Article聽 PubMed聽Google Scholar聽 5.Kudo, I. Murakami, M. Prostaglandin E synthase, a terminal enzyme for prostaglandin E-2 biosynthesis. J. Biochem. Mol. Biol. 38, 633鈥?38 (2005).CAS聽 PubMed聽Google Scholar聽 6.Fahmi, H. MPGES-1 as a novel target for arthritis. Curr. Opin. Rheumatol. 16, 623鈥?27 (2004).CAS聽 Article聽Google Scholar聽 7.Cheng, Y. et al. Cyclooxygenases, microsomal prostaglandin E synthase-1, and cardiovascular function. J. Clin. Invest. 116, 1391鈥?399, https://doi.org/10.1172/jci27540 (2006).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 8.Engblom, D. et al. Microsomal prostaglandin E synthase-1 is the central switch during immune-induced pyresis. Nature Neurosci. 6, 1137鈥?138, https://doi.org/10.1038/nn1137 (2003).CAS聽 Article聽 PubMed聽Google Scholar聽 9.Trebino, C. E. et al. Impaired inflammatory and pain responses in mice lacking an inducible prostaglandin E synthase. Proc. Natl. Acad. Sci. USA 100, 9044鈥?049, https://doi.org/10.1073/pnas.1332766100 (2003).ADS聽 CAS聽 Article聽 PubMed聽Google Scholar聽 10.Schiffler, M. A. et al. Discovery and Characterization of 2-Acylaminoimidazole Microsomal Prostaglandin E Synthase-1 Inhibitors. J. Med. Chem. 59, 194鈥?05, https://doi.org/10.1021/acs.jmedchem.5b01249 (2016).CAS聽 Article聽 PubMed聽Google Scholar聽 11.Hieke, M. et al. Discovery and biological evaluation of a novel class of dual microsomal prostaglandin E2 synthase-1/5-lipoxygenase inhibitors based on 2-[(4,6-diphenethoxypyrimidin-2-yl)thio]hexanoic acid. J. Med. Chem. 54, 4490鈥?507, https://doi.org/10.1021/jm200092b (2011).CAS聽 Article聽 PubMed聽Google Scholar聽 12.Hanke, T. et al. Aminothiazole-featured pirinixic acid derivatives as dual 5-lipoxygenase and microsomal prostaglandin E2 synthase-1 inhibitors with improved potency and efficiency in vivo. J. Med. Chem. 56, 9031鈥?044, https://doi.org/10.1021/jm401557w (2013).ADS聽 CAS聽 Article聽 PubMed聽Google Scholar聽 13.Terracciano, S. et al. Structural Insights for the Optimization of Dihydropyrimidin-2(1H)-one Based mPGES-1 Inhibitors. ACS Med. Chem. Lett. 6, 187鈥?91, https://doi.org/10.1021/ml500433j (2015).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 14.Shiro, T., Kakiguchi, K., Takahashi, H., Nagata, H. Tobe, M. 7-Phenyl-imidazoquinolin-4(5H)-one derivatives as selective and orally available mPGES-1 inhibitors. Bioorg. Med. Chem. 21, 2868鈥?878, https://doi.org/10.1016/j.bmc.2013.03.069 (2013).CAS聽 Article聽 PubMed聽Google Scholar聽 15.Shiro, T., Kakiguchi, K., Takahashi, H., Nagata, H. Tobe, M. Synthesis and biological evaluation of substituted imidazoquinoline derivatives as mPGES-1 inhibitors. Bioorg. Med. Chem. 21, 2068鈥?078, https://doi.org/10.1016/j.bmc.2013.01.018 (2013).CAS聽 Article聽 PubMed聽Google Scholar聽 16.Shiro, T. et al. Synthesis and SAR study of imidazoquinolines as a novel structural class of microsomal prostaglandin E(2) synthase-1 inhibitors. Bioorg. Med. Chem. Lett. 22, 285鈥?88, https://doi.org/10.1016/j.bmcl.2011.11.015 (2012).CAS聽 Article聽 PubMed聽Google Scholar聽 17.Liedtke, A. J. et al. Arylpyrrolizines as inhibitors of microsomal prostaglandin E2 synthase-1 (mPGES-1) or as dual inhibitors of mPGES-1 and 5-lipoxygenase (5-LOX). J. Med. Chem. 52, 4968鈥?972, https://doi.org/10.1021/jm900481c (2009).CAS聽 Article聽 PubMed聽Google Scholar聽 18.Shang, E. et al. Benzo[d]isothiazole 1,1-dioxide derivatives as dual functional inhibitors of 5-lipoxygenase and microsomal prostaglandin E(2) synthase-1. Bioorg. Med. Chem. Lett. 24, 2764鈥?767, https://doi.org/10.1016/j.bmcl.2014.04.006 (2014).CAS聽 Article聽 PubMed聽Google Scholar聽 19.Wu, T. Y. et al. Biarylimidazoles as inhibitors of microsomal prostaglandin E2 synthase-1. Bioorg. Med. Chem. Lett. 20, 6978鈥?982, https://doi.org/10.1016/j.bmcl.2010.09.129 (2010).CAS聽 Article聽 PubMed聽Google Scholar聽 20.Wiegard, A., Hanekamp, W., Griessbach, K., Fabian, J. Lehr, M. Pyrrole alkanoic acid derivatives as nuisance inhibitors of microsomal prostaglandin E2 synthase-1. Eur. J. Med. Chem. 48, 153鈥?63, https://doi.org/10.1016/j.ejmech.2011.12.009 (2012).CAS聽 Article聽Google Scholar聽 21.Chini, M. G. et al. Design and synthesis of a second series of triazole-based compounds as potent dual mPGES-1 and 5-lipoxygenase inhibitors. Eur. J. Med. Chem. 54, 311鈥?23, https://doi.org/10.1016/j.ejmech.2012.05.014 (2012).ADS聽 CAS聽 Article聽 PubMed聽Google Scholar聽 22.Giroux, A. et al. Discovery of disubstituted phenanthrene imidazoles as potent, selective and orally active mPGES-1 inhibitors. Bioorg. Med. Chem. Lett. 19, 5837鈥?841, https://doi.org/10.1016/j.bmcl.2009.08.085 (2009).CAS聽 Article聽 PubMed聽Google Scholar聽 23.Xu, D. et al. MF63 [2-(6-chloro-1H-phenanthro[9,10-d]imidazol-2-yl)-isophthalonitrile], a selective microsomal prostaglandin E synthase-1 inhibitor, relieves pyresis and pain in preclinical models of inflammation. J. Pharmacol. Exp. Ther. 326, 754鈥?63, https://doi.org/10.1124/jpet.108.138776 (2008).CAS聽 Article聽 PubMed聽Google Scholar聽 24.Lee, K. et al. Fragment-based discovery of novel and selective mPGES-1 inhibitors Part 1: identification of sulfonamido-1,2,3-triazole-4,5-dicarboxylic acid. Bioorg. Med. Chem. Lett. 23, 75鈥?0, https://doi.org/10.1016/j.bmcl.2012.11.019 (2013).CAS聽 Article聽 PubMed聽Google Scholar聽 25.Cote, B. et al. Substituted phenanthrene imidazoles as potent, selective, and orally active mPGES-1 inhibitors. Bioorg. Med. Chem. Lett. 17, 6816鈥?820, https://doi.org/10.1016/j.bmcl.2007.10.033 (2007).ADS聽 CAS聽 Article聽 PubMed聽Google Scholar聽 26.Riendeau, D. et al. Inhibitors of the inducible microsomal prostaglandin E2 synthase (mPGES-1) derived from MK-886. Bioorg. Med. Chem. Lett. 15, 3352鈥?355, https://doi.org/10.1016/j.bmcl.2005.05.027 (2005).CAS聽 Article聽 PubMed聽Google Scholar聽 27.Bruno, A. et al. Effects of AF3442 [N-(9-ethyl-9H-carbazol-3-yl)-2-(trifluoromethyl)benzamide], a novel inhibitor of human microsomal prostaglandin E synthase-1, on prostanoid biosynthesis in human monocytes in vitro. Biochem. Pharmacol. 79, 974鈥?81, https://doi.org/10.1016/j.bcp.2009.11.011 (2010).CAS聽 Article聽 PubMed聽Google Scholar聽 28.Koeberle, A. et al. Discovery of benzo[g]indol-3-carboxylates as potent inhibitors of microsomal prostaglandin E(2) synthase-1. Bioorg. Med. Chem. 17, 7924鈥?932, https://doi.org/10.1016/j.bmc.2009.10.025 (2009).CAS聽 Article聽 PubMed聽Google Scholar聽 29.Walker, D. P. et al. Synthesis and biological evaluation of substituted benzoxazoles as inhibitors of mPGES-1: use of a conformation-based hypothesis to facilitate compound design. Bioorg. Med. Chem. Lett. 23, 1120鈥?126, https://doi.org/10.1016/j.bmcl.2012.11.107 (2013).CAS聽 Article聽 PubMed聽Google Scholar聽 30.Wang, J. et al. Selective inducible microsomal prostaglandin E(2) synthase-1 (mPGES-1) inhibitors derived from an oxicam template. Bioorg. Med. Chem. Lett. 20, 1604鈥?609, https://doi.org/10.1016/j.bmcl.2010.01.060 (2010).CAS聽 Article聽 PubMed聽Google Scholar聽 31.Jin, Y. et al. Pharmacodynamic comparison of LY3023703, a novel microsomal prostaglandin e synthase 1 inhibitor, with celecoxib. Clin. Pharmacol. Therap. 99, 274鈥?84, https://doi.org/10.1002/cpt.260 (2016).CAS聽 Article聽Google Scholar聽 32.Ding, K. et al. Structure-based discovery of mPGES-1 inhibitors suitable for preclinical testing in wild-type mice as a new generation of anti-inflammatory drugs. Sci. Rep. 8, 5205, https://doi.org/10.1038/s41598-41018-23482-41594 (2018).ADS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 33.Zhou, Z. et al. Selective inhibitors of human mPGES-1 from structure-based computational screening. Biorg. Med. Chem. Lett. 27, 3739鈥?743 (2017).CAS聽 Article聽Google Scholar聽 34.Ding, K. et al. Design, synthesis, and discovery of 5-((1,3-diphenyl-1H-pyrazol-4-yl)methylene)pyrimidine-2,4,6(1H,3H,5H)-triones and related derivatives as novel inhibitors of mPGES-1. Bioorg. Med. Chem. Lett. 28, 858鈥?62 (2018).CAS聽 Article聽Google Scholar聽 35.Zhou, S. et al. In Silico Observation of the Conformational Opening of the Glutathione-binding Site of Microsomal Prostaglandin E2 Synthase-1. J. Chem. Inf. Model. 59, 3839鈥?845 (2019).CAS聽 Article聽Google Scholar聽 36.Mathieu, M. Enzo Compounds - Knowledge Base: Compound Purity, http://kb.enzolifesciences.com/article/what-is-the-purity-of-the-compounds-included-in-your-screen-well-natural-product-library-bml-2865-611.html (2017).37.Nair, A. B. Jacob, S. A simple practice guide for dose conversion between animals and human. J. Basic Clin. Pharm. 7, 27鈥?1 (2016).Article聽Google Scholar聽 38.Bence, A. K. et al. Phase I pharmacokinetic studies evaluating single and multiple doses of oral GW572016, a dual EGFR-ErbB2 inhibitor, in healthy subjects. Invest. New Drugs 23, 39鈥?9, https://doi.org/10.1023/B:DRUG.0000047104.45929.ea (2005).CAS聽 Article聽 PubMed聽Google Scholar聽 39.Chien, A. J. et al. Phase I Dose-Escalation Study of 5-Day Intermittent Oral Lapatinib Therapy in Patients With Human Epidermal Growth Factor Receptor 2鈥揙verexpressing Breast Cancer. J. Clin. Oncol. 32, 1472鈥?479 (2014).CAS聽 Article聽Google Scholar聽 40.Trebino, C. E. et al. Redirection of Eicosanoid Metabolism in mPGES-1-deficient Macrophages. J. Biol. Chem. 280, 16579鈥?6585 (2005).CAS聽 Article聽Google Scholar聽 41.Kamei, D. et al. Reduced Pain Hypersensitivity and Inflammation in Mice Lacking Microsomal Prostaglandin E synthase-1. J. Biol. Chem. 279, 33684鈥?3695 (2004).CAS聽 Article聽Google Scholar聽 42.Boulet, L. et al. Deletion of Microsomal Prostaglandin E2 (PGE2) synthase-1 Reduces Inducible and Basal PGE2 Production and Alters the Gastric Prostanoid Profile. J. Biol. Chem. 279, 23229鈥?3237 (2004).CAS聽 Article聽Google Scholar聽 43.O鈥橞oyle, N. M. et al. Open Babel: An open chemical toolbox. J. Cheminformatics 3, 33 (2011).Article聽Google Scholar聽 44.Trott, O. Olson, A. J. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 31, 455鈥?61, https://doi.org/10.1002/jcc.21334 (2010).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 45.Case, D. A. et al. AMBER 12, University of California, San Francisco. (2012).46.Andersen, H. C. Rattle: A 鈥渧elocity鈥?version of the shake algorithm for molecular dynamics calculations. J. Comput. Phys. 52, 24鈥?4 (1983).ADS聽 CAS聽 Article聽Google Scholar聽 47.Miller, B. R. 3rd et al. MMPBSA.py: An Efficient Program for End-State Free Energy Calculations. J. Chem. Theo. Comput. 8, 3314鈥?321, https://doi.org/10.1021/ct300418h (2012).CAS聽 Article聽Google Scholar聽 48.Rastelli, G., Degliesposti, G., Del Rio, A. Sgobba, M. Binding estimation after refinement, a new automated procedure for the refinement and rescoring of docked ligands in virtual screening. Chem. Biol. Drug Design 73, 283鈥?86 (2009).CAS聽 Article聽Google Scholar聽 49.Degliesposti, G., Portioli, C., Parenti, M. D. Rastelli, G. BEAR, a novel virtual screening methodology for drug discovery. J. Biomol. Screening 16, 129鈥?33 (2011).CAS聽 Article聽Google Scholar聽 50.Dong, L., Yuan, C., Orlando, B. J., Malkowski, M. G. Smith, W. L. Fatty Acid Binding to the Allosteric Subunit of Cyclooxygenase-2 Relieves a Tonic Inhibition of the Catalytic Subunit. J. Biol. Chem. 291, 25641鈥?5655, https://doi.org/10.1074/jbc.M116.757310 (2016).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 51.Zhou, Z. et al. Selective inhibitors of human mPGES-1 from structure-based computational screening. Bioorg. Med. Chem. Lett. 27, 3739鈥?743, https://doi.org/10.1016/j.bmcl.2017.06.075 (2017).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 52.Hamza, A. et al. Understanding microscopic binding of human microsomal prostaglandin E synthase-1 (mPGES-1) trimer with substrate PGH2 and cofactor GSH: insights from computational alanine scanning and site-directed mutagenesis. J. Phys. Chem. B. 114, 5605鈥?616 (2010).CAS聽 Article聽Google Scholar聽 53.Huang, X. Q. et al. Structural and functional characterization of human microsomal prostaglandin E synthase-1 by computational modeling and site-directed mutagenesis. Bioorg. Med. Chem. 14, 3553鈥?562 (2006).ADS聽 CAS聽 Article聽Google Scholar聽 54.Hamza, A., Zhao, X., Tong, M., Tai, H. H. Zhan, C. G. Novel human mPGES-1 inhibitors identified through structure-based virtual screening. Bioorg. Med. Chem. 19, 6077鈥?086, https://doi.org/10.1016/j.bmc.2011.08.040 (2011).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 55.Edwards, J. C., Sedgwick, A. D. Willoughby, D. A. The formation of a structure with the features of synovial lining by subcutaneous injection of air: an in vivo tissue culture system. J. Pathology 134, 147鈥?56 (1981).CAS聽 Article聽Google Scholar聽 56.Hudachek, S. F. Gustafson, D. L. Physiologically based pharmacokinetic model of lapatinib developed in mice and scaled to humans. J. Pharmacokinet. Pharmacodyn. 40, 157鈥?76, https://doi.org/10.1007/s10928-012-9295-8 (2013).CAS聽 Article聽 PubMed聽 PubMed Central聽Google Scholar聽 Download referencesAcknowledgementsThis work was supported in part by the funding of the Molecular Modeling and Biopharmaceutical Center at the University of Kentucky College of Pharmacy and the National Science Foundation (NSF grant CHE-1111761). The authors also acknowledge the Computer Center at the University of Kentucky for supercomputing time on a Dell Supercomputer Cluster consisting of 388 nodes or 4,816 processors.Author informationAuthor notesThese authors contributed equally: Shuo Zhou, Ziyuan Zhou and Kai Ding.AffiliationsMolecular Modeling and Biopharmaceutical Center, College of Pharmacy, University of Kentucky, 789 South Limestone Street, Lexington, KY, 40536, USAShuo Zhou,聽Ziyuan Zhou,聽Kai Ding,聽Yaxia Yuan,聽Fang Zheng聽 聽Chang-Guo ZhanDepartment of Pharmaceutical Sciences, College of Pharmacy, University of Kentucky, 789 South Limestone Street, Lexington, KY, 40536, USAShuo Zhou,聽Ziyuan Zhou,聽Kai Ding,聽Yaxia Yuan,聽Charles Loftin,聽Fang Zheng聽 聽Chang-Guo ZhanAuthorsShuo ZhouView author publicationsYou can also search for this author in PubMed聽Google ScholarZiyuan ZhouView author publicationsYou can also search for this author in PubMed聽Google ScholarKai DingView author publicationsYou can also search for this author in PubMed聽Google ScholarYaxia YuanView author publicationsYou can also search for this author in PubMed聽Google ScholarCharles LoftinView author publicationsYou can also search for this author in PubMed聽Google ScholarFang ZhengView author publicationsYou can also search for this author in PubMed聽Google ScholarChang-Guo ZhanView author publicationsYou can also search for this author in PubMed聽Google ScholarContributionsS.Z. performed the virtual screening and clinical data mining. Z.Z. did the in vitro activity assays and measured the PGE2 concentrations in the tissue samples from mice. K.D. and C.L. carried out the in vivo experiments. Y.Y. contributed to the computational modeling studies. C.-G.Z. and F.Z. designed the study, analyzed the data, and wrote the manuscript.Corresponding authorsCorrespondence to Fang Zheng or Chang-Guo Zhan.Ethics declarations Competing interests The authors declare no competing interests. Additional informationPublisher鈥檚 note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary Information Supplementary Information.Rights and permissions Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article鈥檚 Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article鈥檚 Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/. Reprints and PermissionsAbout this articleCite this articleZhou, S., Zhou, Z., Ding, K. et al. DREAM-in-CDM Approach and Identification of a New Generation of Anti-inflammatory Drugs Targeting mPGES-1. Sci Rep 10, 10187 (2020). https://doi.org/10.1038/s41598-020-67283-0Download citationReceived: 21 March 2020Accepted: 01 June 2020Published: 23 June 2020DOI: https://doi.org/10.1038/s41598-020-67283-0 CommentsBy submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate. Sign up for the Nature Briefing newsletter 鈥?what matters in science, free to your inbox daily.

本文链接: http://targeting.immuno-online.com/view-41417.html

发布于 : 2024-05-02 阅读()