Chemical Protectors against the Toxic Effects of Paracetamol

Dec 27, 2018 - However, considering physicochemical properties that may affect bioavailability and cell permeability, DHL seems to be the best prospec...
0 downloads 0 Views 920KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 18582−18591

http://pubs.acs.org/journal/acsodf

Chemical Protectors against the Toxic Effects of Paracetamol (Acetaminophen) and Its Meta Analogue: Preventing Protein Arylation Romina Castañeda-Arriaga,† Adriana Pérez-González,‡ and Annia Galano*,† †

Downloaded via 146.185.203.67 on January 1, 2019 at 09:17:13 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, Iztapalapa, C. P., 09340 México D. F., Mexico ‡ CONACYTUniversidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina. Iztapalapa, C. P., 09340 México D. F., Mexico S Supporting Information *

ABSTRACT: The potential role of nine thiols as chemical protectors against the toxicity of paracetamol (acetaminophen, APAP) and its meta analogue N-acetyl-m-aminophenol (AMAP) was investigated using the density functional theory. They are glutathione (GSH), N-acetylcysteine (NAC), NAC amide (NACA), tiopronine (TPR), dihydrolipoic acid (DHL), 6-mercaptopurine (6MP), 6-thioguanine (6TG), 2,3-dimercaprol (DMC), and D-penicillamine (PNA). The investigation was focused on the toxic effects derived from protein arylation at cysteine residues, induced by the quinone imines formed from APAP and AMAP. On the basis of both thermochemical and kinetic considerations, with the exceptions of 6MP and 6TG, the investigated thiols may be useful in protecting the chemical integrity of cysteine residues in proteins from arylation by quinone imines. The order of efficiency for that purpose is predicted to be NAC > GSH > TPR > NACA > DMC > DHL. However, considering physicochemical properties that may affect bioavailability and cell permeability, DHL seems to be the best prospect to be orally supplied.



INTRODUCTION N-Acetyl-p-aminophenol (i.e., paracetamol, acetaminophen, or APAP; Scheme 1), is a widely used analgesic and antipyretic.

Scheme 2. Structure of NAPQI, PBQI, and Related Species, with Site Numbering

Scheme 1. Structures of Paracetamol (APAP), Its Meta Analogue (AMAP)

Although safe at therapeutic doses, APAP overdoses can be toxic. Such toxicity is considered the most frequent cause of drug-induced liver injuries1−3 and is also known to affect other organs.4−10 N-Acetyl-p-benzoquinone imine (NAPQI, Scheme 2) is a metabolite of APAP, which has been identified as responsible for the APAP hepatotoxicity.3,11 At APAP therapeutic doses, NAPQI is properly detoxified by glutathione (GSH). On the contrary, when APAP is consumed in excess, NAPQI accumulates and GSH is insufficient for preventing the © 2018 American Chemical Society

associated injuries.2,3,11,12 Chemically speaking, such injures are characterized by the formation of NAPQI adducts with the thiol group in cysteine residues, which eventually result in the Received: October 25, 2018 Accepted: December 13, 2018 Published: December 27, 2018 18582

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega

Article

production of the oxidative species.13,14 The adduct formation mainly involves covalent bonding between the thiol group in cysteine residues and the ortho position, with respect to the keto group, in NAPQI.15 In addition, p-benzoquinone imine (PBQI, Scheme 2), which can be produced by deacetylation of NAPQI,16 has been proposed as a potential contributor to the APAP toxicity.17 The paracetamol meta analogue, N-acetyl-m-aminophenol (AMAP, Scheme 1), has been suggested as a safer alternative to APAP.18,19 However, there is also evidence indicating otherwise.15,20,21 AMAP metabolites (Scheme 2), the yield of which is similar to that of APAP, have also been found to form adducts with cysteine residues, by binding to the thiol group.15 Therefore, it seems that different approaches for preventing APAP (and AMAP) toxicity are required. Chemical protectors are likely candidates for that purpose. As previously mentioned, endogenously produced GSH is known to offer protection against the toxic effects of APAP. GSH is a tripeptide containing cysteine as the central residue (Scheme 3). Thus, its nonenzymatic protection involves its

amide group replacing the carboxyl group in NAC, which increases lipophilicity and facilitates crossing cell membranes. Therefore, lower doses of NACA (compared to NAC) are required to counteract APAP poisoning, which may prevent adverse side effects.52 Some studies in animal models have shown that NACA is more efficient than NAC against APAPinduced toxicity because of the higher bioavailability of NACA, although both have the same ways of action.53 In addition, NACA has been proved to have antioxidant effects, which are capable of scavenging ROS, and avoid oxidative stress.52,55 Thus, it is expected to prevent the damaging effects of the oxidative species arising from the APAP toxicity. The presence of the thiol group seems to be the key structural feature in APAP antidotes. This is logical because this group would bind to NAPQI preventing protein arylation at cysteine residues. In addition, thiols are usually good antioxidants known to efficiently deactivate ROS. Therefore, the potential role of several thiols as chemical protectors against the toxic effects of APAP and AMAP has been investigated in this work. To that purpose, protection mechanism (ii), that is, direct binding to NAPQI and analogues, has been considered. The corresponding thermochemical and kinetic data were compared with those of GSH, NACA, and NAC. On the basis of these comparisons, some of the investigated compounds are proposed as promising alternatives for the use of NAC in the treatment of APAP poisoning.

Scheme 3. Structures of the Thiol Investigated in This Work



INVESTIGATED COMPOUNDS Nine thiols were chosen for this investigation (Scheme 3). They are tiopronine (TPR), dihydrolipoic acid (DHL), 6mercaptopurine (6MP), 6-thioguanine (6TG), 2,3-dimercaprol (DMC), D-penicillamine (PNA), NAC, NACA, and GSH. Although 6MP and 6TG have two tautomeric forms (Figure S1, Supporting Information), those included in the present investigation are the ones with the thiol group. As discussed in the Introduction section, NAC, NACA, and GSH are already known to counteract the toxic effects of APAP. To the best of our knowledge, the other investigated compounds have been used, so far, for other purposes. Among them, DHL is probably the most widely used thiol. Its therapeutic action has been frequently associated with its antioxidant properties, which involve free-radical scavenging and protein-repairing capabilities.56 It has been reported to offer neuroprotection, in vitro, against oxidative stress.57 It was also found that 1 month treatment with DHL helps mitigating the pain symptoms typical of several stress oxidative-dependent diseases.58 The antioxidant, antiinflammatory, and immunomodulatory activities of DHL have been identified as very helpful in preventing miscarriage and preterm delivery in pregnant women.59 Current studies support using DHL in the auxiliary treatment of many diseases including cancer, acquired immune deficiency syndrome, diabetes, hypertension, and inflammation as well as cardiovascular, autoimmune, hepatic, and neurodegenerative disorders.60−65 DHL improves neurofunction and alleviated inflammation in early brain injury, after subarachnoid hemorrhage.66 It also has inhibitory effects on the viability and proliferation of breast cancer cells.67 On the other hand, TPR has been used to treat cystinuria,68−70 rheumatoid arthritis,71−75 and liver damage.76,77 It also has antioxidant effects, as free-radical scavenger, metal chelator, and inhibitor of the formation of OH radicals via Fenton reactions.78,79 6MP has been found to be effective

role as a sacrifice target that forms adducts with NAPQI,22 preventing protein binding.23 It has been estimated that, in vivo, the formation of such adducts takes place at a rate of 3.4 × 104 M−1 s−1 nonenzymatically and at a rate of 2.0 × 107 M−1 s−1 when catalyzed by GSH S-transferases.24 In addition, GSH is well known for its antioxidant activity,25−33 which is expected to help counteracting the effects of the oxidative species34 formed as a consequence of the APAP toxicity. Other chemicals containing the thiol group have been tested, or used, as antidotes against APAP poisoning. Some of initially used chemicals in this context were cysteine, methionine, cysteamine, and N-acetylcysteine (NAC),35−38 the latter being the most widely used in this context. Several mechanisms have been proposed to contribute to the NAC protection against the hepatotoxicity of APAP: (i) by providing cysteine for GSH synthesis;39,40 (ii) by forming adducts with NAPQI;24,41 (iii) by reacting with reactive oxygen species (ROS);15 (iv) by reducing NAPQI back to APAP.24,42,43 However, some adverse effects have been associated with NAC administration as an antidote of APAP overdose including anaphylactoid reactions, nausea, vomiting, headache, and renal failure.44−51 In addition, because of its poor bioavailability (the carboxyl group deprotonates at physiological pH, making NAC negatively charged),52 high doses and/or long treatment times of NAC are required. NAC amide (NACA) has emerged as a promising alternative to NAC for treating APAP-induced toxicity.53,54 NACA has an 18583

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega

Article

Table 1. Acid Constants, Expressed as pKa Values, of the Thiols Investigated in the Present Work; the Molar Fractions (Mf) of the Neutral (HnA), Monoanionic (Hn−1A−), and Dianionic (Hn−2A2−) Species at pH = 7.4 pKa1 NAC NACA TPR DHL GSH 6MP 6TG PNA DMC

b

3.11 9.12a 3.59b 4.85c 3.59b 2.25a 2.71a 1.90b 8.63b

pKa2

pKa3 b

9.58

8.87b 10.7c 8.75b 4.19a 5.05a 7.96b 10.65b

10.67b

M

f (HnA)

0.0001 0.9813 0.0001 0.0028 0.0001 0.0006 0.0040 0.7840 0.9444

M

f (Hn−1A−) 0.9934 0.0187 0.9671 0.9967 0.9571 0.9994 0.9960 0.2159 0.0556

M

f (Hn−2A2−) 0.0066 0.0328 0.0005 0.0428

0.0001 GSH > TPR > NACA > DMC > DHL. However, considering physicochemical properties that may affect bioavailability and cell permeability, DHL seems to be the best prospect to be orally supplied.

quinone imines is step 1, that is, the initial addition. It should be kept in mind, though, that pH influences step 3. Thus, this step might become endergonic under high acid conditions. Other Considerations. On the basis of both thermochemical and kinetic considerations, most of the investigated thiols were identified as promising candidates for protecting the chemical integrity of cysteine residues in proteins from arylation by quinone imines. The order of efficiency for that purpose would be NAC > GSH > TPR > NACA > DMC > DHL. However, there are other important factors to consider that may affect bioavailability and cell permeability. Therefore, several physicochemical parameters usually used as indicators for absorption, distribution, metabolism, and excretion (ADME) properties were estimated for all the investigated thiols. These parameters were chosen to check if the thiols satisfy Lipinski’s,116 Ghose’s,117 and Veber’s118 criteria and are reported in Table S13 (Supporting Information). As previously mentioned, the partition coefficient octanol/ water (log P) is considered problematic for NAC (log P = −2.44), which was one of the reasons of using NACA (log P = −0.94) instead. Considering Lipinski’s and Ghose’s rules together, the desired value of log P should range from −0.4 to 5.0. Among the thiols proposed as the best protectors against protein arylation, only DHL (log P = 2.01) and DMC (log P = 0.31) fulfill this requirement. The log P value for TPR (log P = −1.56) is between those of NAC and NACA, while that of GSH (log P = −4.97) is the one differing the most for the target values. Regarding the number of H bond acceptors (HBA), H bond donors (HBD), and rotatable bonds, all the investigated thiols are within the desirable ranges except for GSH, with HBA = 6. All the thiols comply with Lipinski’s rule for molecular weight (MW), although that of DMC (MW = 124.23) is to some extent below Ghose’s low threshold (160). Ghose’s rules for the number of nonhydrogen atoms (XAt) and molar refractivity (MR) are the ones for which most of the investigated thiols are outside the desirable ranges (20 ≤ XAt ≤ 70, 40 ≤ MR ≤ 130). They all have less than 20 XAt, with the exception of GSH. Their MR values are slightly lower than 40 for all of them but GSH and DHL. Thus, considering the ADME properties estimated here, DHL seems to be the best prospect to be orally supplied for preventing paracetamol toxicity associated with protein arylation. The other thiols might present bioavailability and/ or cell permeability issues. However, Lipinski’s, Ghose’s, and Veber’s criteria are only guiding rules, and they have been estimated here through structure−activity relationships. Therefore, this point should be further explored, preferably by experimental investigations.



COMPUTATIONAL DETAILS Geometry optimizations and frequency calculations were performed within the framework of the density functional theory, with the Gaussian 09 package of programs.119 In particular, the M05-2X functional120 was chosen because it is recommended for kinetic calculations.120,121 It was used in conjunction with the 6-31+G(d,p) basis set and the solvation model based on density (SMD),122 using water as solvent to mimic aqueous environments. SMD is considered to be a universal solvation model, applicable to any charged or uncharged solute in any solvent or liquid medium.122 Local minima and saddle points were identified by the number of imaginary frequencies (0 and 1, respectively). Thermodynamic corrections at 298.15 K were included in the calculation of relative energies, which correspond to the 1 M standard state. The rate constants (k) were calculated using the same standard state and the conventional transition-state theory123−125 with harmonic vibrational frequencies and Eckart tunneling.126,127 Kinetic calculations were not performed for the endergonic reaction pathways because they are expected to be reversible to a significant extent. The abundance of acid− base species at physiological pH was considered in the calculation of the rate constants. The overall rate coefficients were obtained as the sum of the rate coefficients corresponding to the reaction pathways identified as viable. More details on how to apply this procedure can be found elsewhere.15,128 Calculated rate constants (k) close to the diffusion limit were corrected using the Collins−Kimball theory,129 the steady-state Smoluchowski rate constant for an irreversible bimolecular diffusion-controlled reaction,130 and the Stokes−Einstein131,132 approaches. This strategy has been previously validated.128 Acid constants, expressed as pKa, were calculated using the fitted parameters approach.133,134 It has been demonstrated that this approach produces deviations from experiments that are systematically lower than 0.5 pKa units, in terms of mean unsigned errors. This approach uses experimental pKa values to obtain two parameters (m and C0) from linear fittings, which are then used to calculate the pKa of the target molecule. The parameters used here have been previously proposed and validated.133



CONCLUSIONS A previously proposed mechanism for the reactions between thiols and quinone imine was used to investigate the potential role of nine thiols as chemical protectors against the toxicity of paracetamol (acetaminophen) and its meta analogue. The investigated protection is intended to prevent protein arylation at cysteine residues. The first step (addition of quinone imines to the thiol group) was identified as the rate-determining step and the crucial one for the target reaction. It was found to involve only the deprotonate fractions (i.e., the thiolate anions). Thus, the pKa of the thiols, and their relationship with the pH of the environment, are expected to influence the efficiency of thiols



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b02943. Gibbs free energies and activation energies of all the investigated reaction steps, ADME properties of the investigated thiols, structures of 6MP and 6TG 18587

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega



Article

mechanisms, analogues and protective approaches. Crit. Rev. Toxicol. 2008, 31, 55−138. (12) McGill, M. R.; Lebofsky, M.; Norris, H.-R. K.; Slawson, M. H.; Bajt, M. L.; Xie, Y.; Williams, C. D.; Wilkins, D. G.; Rollins, D. E.; Jaeschke, H. Plasma and liver acetaminophen-protein adduct levels in mice after acetaminophen treatment: Dose-response, mechanisms, and clinical implications. Toxicol. Appl. Pharmacol. 2013, 269, 240− 249. (13) Bajt, M. L.; Knight, T. R.; Lemasters, J. J.; Jaeschke, H. Acetaminophen-Induced Oxidant Stress and Cell Injury in Cultured Mouse Hepatocytes: Protection by N-Acetyl Cysteine. Toxicol. Sci. 2004, 80, 343−349. (14) Jaeschke, H. Glutathione disulfide formation and oxidant stress during acetaminophen-induced hepatotoxicity in mice in vivo: The protective effect of allopurinol. J. Pharmacol. Exp. Ther. 1990, 9, 935− 941. (15) Castañeda-Arriaga, R.; Galano, A. Exploring Chemical Routes Relevant to the Toxicity of Paracetamol and Its meta-Analogue at a Molecular Level. Chem. Res. Toxicol. 2017, 30, 1286−1301. (16) Högestätt, E. D.; Jönsson, B. A. G.; Ermund, A.; Andersson, D. A.; Björk, H.; Alexander, J. P.; Cravatt, B. F.; Basbaum, A. I.; Zygmunt, P. M. Conversion of Acetaminophen to the BioactiveNAcylphenolamine AM404 via Fatty Acid Amide Hydrolase-dependent Arachidonic Acid Conjugation in the Nervous System. J. Biol. Chem. 2005, 280, 31405−31412. (17) Fischer, V.; West, P. R.; Nelson, S. D.; Harvison, P. J.; Mason, R. P. Formation of 4-aminophenoxyl free radical from the acetaminophen metabolite N-acetyl-p-benzoquinone imine. J. Biol. Chem. 1985, 260, 11446−11450. (18) Kyriakides, M.; Maitre, L.; Stamper, B. D.; Mohar, I.; Kavanagh, T. J.; Foster, J.; Wilson, I. D.; Holmes, E.; Nelson, S. D.; Coen, M. Comparative metabonomic analysis of hepatotoxicity induced by acetaminophen and its less toxic meta-isomer. Arch. Toxicol. 2016, 90, 3073−3085. (19) Stamper, B. D.; Bammler, T. K.; Beyer, R. P.; Farin, F. M.; Nelson, S. D. Differential Regulation of Mitogen-Activated Protein Kinase Pathways by Acetaminophen and Its Nonhepatotoxic Regioisomer 3′-Hydroxyacetanilide in TAMH Cells. Toxicol. Sci. 2010, 116, 164−173. (20) Hadi, M.; Dragovic, S.; van Swelm, R.; Herpers, B.; van de Water, B.; Russel, F. G. M.; Commandeur, J. N. M.; Groothuis, G. M. M. AMAP, the alleged non-toxic isomer of acetaminophen, is toxic in rat and human liver. Arch. Toxicol. 2012, 87, 155−165. (21) Xie, Y.; McGill, M. R.; Du, K.; Dorko, K.; Kumer, S. C.; Schmitt, T. M.; Ding, W.-X.; Jaeschke, H. Mitochondrial protein adducts formation and mitochondrial dysfunction during N-acetyl-maminophenol (AMAP)-induced hepatotoxicity in primary human hepatocytes. Toxicol. Appl. Pharmacol. 2015, 289, 213−222. (22) Rosen, G. M.; Rauckman, E. J.; Ellington, S. P.; Dahlin, D. C.; Christie, J. L.; Nelson, S. D. Reduction and glutathione conjugation reactions of N-acetyl-p-benzoquinone imine and two dimethylated analogues. Mol. Pharmacol. 1984, 25, 151−157. (23) Dahlin, D. C.; Miwa, G. T.; Lu, A. Y.; Nelson, S. D. N-acetyl-pbenzoquinone imine: A cytochrome P-450-mediated oxidation product of acetaminophen. Isotopenpraxis 1984, 81, 1327−1331. (24) Coles, B.; Wilson, I.; Wardman, P.; Hinson, J. A.; Nelson, S. D.; Ketterer, B. The spontaneous and enzymatic reaction of N-acetyl-pbenzoquinonimine with glutathione: A stopped-flow kinetic study. Arch. Biochem. Biophys. 1988, 264, 253−260. (25) Jan, A. T.; Ali, A.; Haq, Q. M. R. Glutathione as an antioxidant in inorganic mercury induced nephrotoxicity. J. Postgrad. Med. 2011, 57, 72−77. (26) Meister, A. Glutathione-ascorbic acid antioxidant system in animals. J. Biol. Chem. 1994, 269, 9397−400. (27) Mugesh, G. Glutathione peroxidase activity of ebselen and its analogues: Some insights into the complex chemical mechanisms underlying the antioxidant activity. Curr. Chem. Biol. 2013, 7, 47−56.

tautomers, and optimized structures of the transition states (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected], [email protected]. ORCID

Annia Galano: 0000-0002-1470-3060 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully acknowledge the Laboratorio de Visualización y Cómputo Paralelo at Universidad Autónoma MetropolitanaIztapalapa. A.P.-G. Acknowledges the economic support of the Program of CátedrasCONACYT from CONACyTUAMI (2015−2025), ID-Investigador 435. R.C.-A. acknowledges the financial support through a CONACyT postdoctoral fellowship, through project IFC-2016/1828.



REFERENCES

(1) Lee, W. M. Recent developments in acute liver failure. Best Pract. Res., Clin. Gastroenterol. 2012, 26, 3−16. (2) McGill, M. R.; Sharpe, M. R.; Williams, C. D.; Taha, M.; Curry, S. C.; Jaeschke, H. The mechanism underlying acetaminopheninduced hepatotoxicity in humans and mice involves mitochondrial damage and nuclear DNA fragmentation. J. Clin. Invest. 2012, 122, 1574−1583. (3) Graham, G. G.; Davies, M. J.; Day, R. O.; Mohamudally, A.; Scott, K. F. The modern pharmacology of paracetamol: Therapeutic actions, mechanism of action, metabolism, toxicity and recent pharmacological findings. Inflammopharmacology 2013, 21, 201−232. (4) Kalinec, G. M.; Thein, P.; Parsa, A.; Yorgason, J.; Luxford, W.; Urrutia, R.; Kalinec, F. Acetaminophen and NAPQI are toxic to auditory cells via oxidative and endoplasmic reticulum stressdependent pathways. Hear. Res. 2014, 313, 26−37. (5) Kristensen, D. M.; Hass, U.; Lesné, L.; Lottrup, G.; Jacobsen, P. R.; Desdoits-Lethimonier, C.; Boberg, J.; Petersen, J. H.; Toppari, J.; Jensen, T. K.; Brunak, S.; Skakkebæk, N. E.; Nellemann, C.; Main, K. M.; Jégou, B.; Leffers, H. Intrauterine exposure to mild analgesics is a risk factor for development of male reproductive disorders in human and rat. Hum. Reprod. 2010, 26, 235−244. (6) Watanabe, H.; Kamiyama, T.; Sasaki, S.; Kobayashi, K.; Fukuda, K.; Miyake, Y.; Aruga, T.; Sueki, H. Toxic epidermal necrolysis caused by acetaminophen featuring almost 100% skin detachment: Acetaminophen is associated with a risk of severe cutaneous adverse reactions. J. Dermatol. 2016, 43, 321−324. (7) Pena, M. Á .; Pérez, S.; Concepción Zazo, M.; Alcalá, P. J.; Cuello, J. D.; Zapater, P.; Reig, R. A case of Toxic Epidermal Necrolysis secondary to Acetaminophen in a Child. Curr. Drug Saf. 2016, 11, 99−101. (8) Ghanem, C. I.; Pérez, M. J.; Manautou, J. E.; Mottino, A. D. Acetaminophen from liver to brain: New insights into drug pharmacological action and toxicity. Pharmacol Res 2016, 109, 119−131. (9) McGill, M. R.; Kennon-McGill, S.; Durham, D.; Jaeschke, H. Hearing, reactive metabolite formation, and oxidative stress in cochleae after a single acute overdose of acetaminophen: anin vivostudy. Toxicol. Mech. Methods 2016, 26, 104−111. (10) El-Shafey, M. M.; Abd-Allah, G. M.; Mohamadin, A. M.; Harisa, G. I.; Mariee, A. D. Quercetin protects against acetaminopheninduced hepatorenal toxicity by reducing reactive oxygen and nitrogen species. Pathophysiology 2015, 22, 49−55. (11) Bessems, J. G. M.; Vermeulen, N. P. E. Paracetamol (acetaminophen)-induced toxicity: Molecular and biochemical 18588

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega

Article

acetaminophen overdose. Fundam. Clin. Pharmacol. 2011, 25, 405− 410. (49) Mallet, P.; Mourdi, N.; Dubus, J.-C.; Bavoux, F.; BoyerGervoise, M.-J.; Jean-Pastor, M.-J.; Chalumeau, M. Respiratory paradoxical adverse drug reactions associated with acetylcysteine and carbocysteine systemic use in paediatric patients: A national survey. PLoS One 2011, 6, No. e22792. (50) Sandilands, E. A.; Bateman, D. N. Adverse reactions associated with acetylcysteine. Clin. Toxicol. 2009, 47, 81−88. (51) Pakravan, N.; Waring, W. S.; Sharma, S.; Ludlam, C.; Megson, I.; Bateman, D. N. Risk factors and mechanisms of anaphylactoid reactions to acetylcysteine in acetaminophen overdose. Clin. Toxicol. 2008, 46, 697−702. (52) Ates, B.; Abraham, L.; Ercal, N. Antioxidant and free radical scavenging properties of N-acetylcysteine amide (NACA) and comparison with N-acetylcysteine (NAC). Free Radic. Res. 2009, 42, 372−377. (53) Khayyat, A.; Tobwala, S.; Hart, M.; Ercal, N. N -acetylcysteine amide, a promising antidote for acetaminophen toxicity. Toxicol. Lett. 2016, 241, 133−142. (54) Sunitha, K.; Hemshekhar, M.; Thushara, R. M.; Santhosh, M. S.; Yariswamy, M.; Kemparaju, K.; Girish, K. S. N-Acetylcysteine amide: A derivative to fulfill the promises of N-Acetylcysteine. Free Radic. Res. 2013, 47, 357−367. (55) Galano, A. Mechanism and kinetics of the hydroxyl and hydroperoxyl radical scavenging activity of N-acetylcysteine amide. Theor. Chem. Acc. 2011, 130, 51−60. (56) Castañeda-Arriaga, R.; Alvarez-Idaboy, J. R. Lipoic acid and dihydrolipoic acid. A comprehensive theoretical study of their antioxidant activity supported by available experimental kinetic data. J. Chem. Inf. Model. 2014, 54, 1642−1652. (57) Koenig, M. L.; Meyerhoff, J. L. In vitro neuroprotection against oxidative stress by pre-treatment with a combination of dihydrolipoic acid and phenyl-butyl nitrones. Neurotox. Res. 2003, 5, 265−272. (58) Vigna, L.; Solimena, G.; Bamonti, F.; Arcaro, M.; Fenoglio, C.; Oldoni, E.; Dellanoce, C.; Rossi, P.; Gori, F.; Figliola, R.; Cighetti, G. Effects of 1-month R-α-lipoic acid supplementation on humans oxidative status: A pilot study. Prog. Nutr. 2017, 19, 14−25. (59) Parente, E.; Colannino, G.; Picconi, O.; Monastra, G. Safety of oral alpha-lipoic acid treatment in pregnant women: a retrospective observational study. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 4219− 4227. (60) Gorąca, A.; Huk-Kolega, H.; Piechota, A.; Kleniewska, P.; Ciejka, E.; Skibska, B. Lipoic acid - Biological activity and therapeutic potential. Pharmacol. Rep. 2011, 63, 849−858. (61) Shay, K. P.; Moreau, R. F.; Smith, E. J.; Smith, A. R.; Hagen, T. M. Alpha-lipoic acid as a dietary supplement: Molecular mechanisms and therapeutic potential. Biochim. Biophys. Acta 2009, 1790, 1149− 1160. (62) Rochette, L.; Ghibu, S.; Muresan, A.; Vergely, C. Alpha-lipoic acid: Molecular mechanisms and therapeutic potential in diabetes. Can. J. Physiol. Pharmacol. 2015, 93, 1021−1027. (63) Gomes, M.; Negrato, C. Alpha-lipoic acid as a pleiotropic compound with potential therapeutic use in diabetes and other chronic diseases. Diabetol. Metab. Syndrome 2014, 6, 80. (64) Coletta, C.; Módis, K.; Szczesny, B.; Brunyánszki, A.; Oláh, G.; Rios, E. C. S.; Yanagi, K.; Ahmad, A.; Papapetropoulos, A.; Szabo, C. Regulation of Vascular Tone, Angiogenesis and Cellular Bioenergetics by the 3-Mercaptopyruvate Sulfurtransferase/H2S Pathway: Functional Impairment by Hyperglycemia and Restoration by DL-alphaLipoic Acid. Mol. Med. 2015, 47, S44−S45. (65) Foo, N.-P.; Lin, S.-H.; Lee, Y.-H.; Wu, M.-J.; Wang, Y.-J. αLipoic acid inhibits liver fibrosis through the attenuation of ROStriggered signaling in hepatic stellate cells activated by PDGF and TGF-β. Toxicology 2011, 282, 39−46. (66) Zhou, K.; Enkhjargal, B.; Xie, Z.; Sun, C.; Wu, L.; Malaguit, J.; Chen, S.; Tang, J.; Zhang, J.; Zhang, J. H. Dihydrolipoic acid inhibits lysosomal rupture and NLRP3 through lysosome-associated membrane Protein-1/Calcium/Calmodulin-Dependent Protein Kinase II/

(28) Sciuto, A. M. Antioxidant properties of glutathione and its role in tissue protection. Oxidants, Antioxidants, and Free Radicals; CRC Press, 2017; pp 171−192. (29) Stankov, K.; Kovacevic, Z. Oxidative stress and glutathione Associated antioxidative systems in malignant cells. Arch. Oncol. 1998, 6, 109−113. (30) van Zandwijk, N. N-Acetylcysteine (NAC) and glutathione (GSH): Antioxidant and chemopreventive properties, with special reference to lung cancer. J. Cell. Biochem. 1995, 59, 24−32. (31) Marí, M.; Morales, A.; Colell, A.; García-Ruiz, C.; FernándezCheca, J. C. Mitochondrial glutathione, a key survival antioxidant. Antioxid. Redox Signaling 2009, 11, 2685−2700. (32) Galano, A.; Alvarez-Idaboy, J. R. Glutathione: Mechanism and kinetics of its non-enzymatic defense action against free radicals. RSC Adv. 2011, 1, 1763−1771. (33) Alvarez-Idaboy, J. R.; Galano, A. On the chemical repair of DNA radicals by glutathione: Hydrogen vs electron transfer. J. Phys. Chem. B 2012, 116, 9316−9325. (34) Galano, A.; Mazzone, G.; Alvarez-Diduk, R.; Marino, T.; Alvarez-Idaboy, J. R.; Russo, N. Food Antioxidants: Chemical Insights at the Molecular Level. Annu. Rev. Food Sci. Technol. 2016, 7, 335− 352. (35) Prescott, L. F.; Sutherland, G. R.; Park, J.; Smith, I. J.; Proudfoot, A. T. Cysteamine, methionine, and penicillamine in the treatment of paracetamol poisoning. Lancet 1976, 308, 109−113. (36) Smilkstein, M. J.; Knapp, G. L.; Kulig, K. W.; Rumack, B. H. Efficacy of Oral N-Acetylcysteine in the Treatment of Acetaminophen Overdose. N. Engl. J. Med. 1988, 319, 1557−1562. (37) Prescott, L. F.; Ballantyne, A.; Proudfoot, A. T.; Park, J.; Adriaenssens, P. Treatment of paracetamol (acetaminophen) poisoning with n-acetylcysteine. Lancet 1977, 310, 432−434. (38) Prescott, L. F.; Illingworth, R. N.; Critchley, J. A.; Stewart, M. J.; Adam, R. D.; Proudfoot, A. T. Intravenous N-acetylcystine: the treatment of choice for paracetamol poisoning. Br. Med. J. 1979, 2, 1097−1100. (39) Corcoran, G. B.; Racz, W. J.; Smith, C. V.; Mitchell, J. R. Effects of N-acetylcysteine on acetaminophen covalent binding and hepatic necrosis in mice. J. Pharmacol. Exp. Ther. 1985, 232, 864−872. (40) Corcoran, G. B.; Todd, E. L.; Racz, W. J.; Hughes, H.; Smith, C. V.; Mitchell, J. R. Effects of N-acetylcysteine on the disposition and metabolism of acetaminophen in mice. J. Pharmacol. Exp. Ther. 1985, 232, 857−63. (41) Buckpitt, A. R.; Rollins, D. E.; Mitchell, J. R. Varying effects of sulfhydryl nucleophiles on acetaminophen oxidation and sulfhydryl adduct formation. Biochem. Pharmacol. 1979, 28, 2941−2946. (42) Corcoran, G. B.; Mitchell, J. R.; Vaishnav, Y.; Horning, E. C. Evidence that acetaminophen and N-hydroxyacetaminophen form a common arylating intermediate, N-acetyl-p-benzoquinoneimine. Mol. Pharmacol. 1980, 18, 536−542. (43) Lauterburg, B. H.; Corcoran, G. B.; Mitchell, J. R. Mechanism of action of N-acetylcysteine in the protection against the hepatotoxicity of acetaminophen in rats in vivo. J. Clin. Invest. 1983, 71, 980−991. (44) Koppen, A.; van Riel, A.; de Vries, I.; Meulenbelt, J. Recommendations for the paracetamol treatment nomogram and side effects of N-acetylcysteine. Neth. J. Med. 2014, 72, 251−257. (45) Mahmoudi, G. A.; Astaraki, P.; Mohtashami, A. Z.; Ahadi, M. N-acetylcysteine overdose after acetaminophen poisoning. Int. Med. Case Rep. J. 2015, 8, 65−69. (46) Wong, A.; Graudins, A. N-acetylcysteine regimens for paracetamol overdose: Time for a change? Emerg. Med. Australasia 2016, 28, 749−751. (47) Yamamoto, T.; Spencer, T.; Dargan, P. I.; Wood, D. M. Incidence and management of N-acetylcysteine-related anaphylactoid reactions during the management of acute paracetamol overdose. Eur. J. Emerg. Med. 2014, 21, 57−60. (48) Zyoud, S. H.; Awang, R.; Syed Sulaiman, S. A.; Al-Jabi, S. W. Nacetylcysteine-induced headache in hospitalized patients with acute 18589

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega

Article

TAK1 pathways after subarachnoid hemorrhage in rat. Stroke 2018, 49, 175−183. (67) Kuban-Jankowska, A.; Gorska-Ponikowska, M.; Wozniak, M. Lipoic acid decreases the viability of breast cancer cells and activity of PTP1B and SHP2. Anticancer Res. 2017, 37, 2893−2898. (68) Carlsson, M. S.; Denneberg, T.; Emanuelsson, B.-M.; Kågedal, B.; Lindgren, S. Steady-State Pharmacokinetics of 2-Mercaptopropionylglycine (Tiopronin) in Healthy Volunteers and Patients with Cystinuria. Drug Invest. 2012, 7, 41−51. (69) Lindell, Å.; Denneberg, T.; Jeppsson, J.-O. Urinary excretion of free cystine and the tiopronin-cysteine-mixed disulfide during longterm tiopronin treatment of cystinuria. Nephron 1995, 71, 328−342. (70) Tasic, V.; Lozanovski, V. J.; Ristoska-Bojkovska, N.; Sahpazova, E.; Gucev, Z. Nephrotic syndrome occurring during tiopronin treatment for cystinuria. Eur. J. Pediatr. 2010, 170, 247−249. (71) Albouy, R.; Bendaoud, B.; Casburn-Budd, R.; Chicault, P.; Le Goff, P.; Paolozzi, L.; Peron, A.; Thorel, J. B.; Tirel, H.; Youinou, P. Tiopronin-induced reduction of rheumatoid factor functional affinity. Rev. Rhum. Mal. Osteoartic. 1991, 58, 43−49S. (72) Fauquert, P.; Delecoeuillerie, G.; Lelong, A.; Le Goff, P.; Youinou, P. Phenotypic analysis of circulating T lymphocytes from rheumatoid arthritis patients treated with tiopronine. Int. J. Immunopathol. Pharmacol. 1989, 2, 173−179. (73) Lelong, A.; Delecoeuillerie, G.; Fauquert, P.; Le Goff, P.; Youinou, P. Tiopronine modifies polymorphonuclear cell functions in patients with rheumatoid arthritis. Int. J. Immunother. 1989, 5, 123− 127. (74) Pasero, G.; Pellegrini, P.; Ciompi, M. L.; Colamussi, V.; Barbieri, P.; Mazzoni, M. R. Tiopronine: new basic treatment for rheumatoid arthritis. Rev. Rhum. Mal. Osteoartic. 1980, 47, 163−168. (75) Youinou, P.; Keusch, J.; Casburn-Budd, R.; Paolozzi, L.; Lamour, A.; Le Goff, P.; Bendaoud, B. Tiopronine-induced reduction of rheumatoid factor functional affinity and asialylated IgG in patients with rheumatoid arthritis. Clin. Exp. Rheumatol. 1992, 10, 333−338. (76) Feng, Y. G.; Guo, X. S.; Yan, X. Y.; Wang, C. H.; Du, J.; Zhao, Z.; Pan, M.; Li, Y. Studies on tiopronin in the treatment of hepatic damage induced by antipsychotics. Chin. Pharm. J. 2003, 38, 64−66. (77) Tang, M. C.; Cheng, L.; Qiu, L.; Jia, R. G.; Sun, R. Q.; Wang, X. P.; Hu, G. Y.; Zhao, Y. Efficacy of Tiopronin in treatment of severe non-alcoholic fatty liver disease. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 160−164. (78) Ghibu, S.; Richard, C.; Vergely, C.; Zeller, M.; Cottin, Y.; Rochette, L. Antioxidant properties of an endogenous thiol: Alphalipoic acid, useful in the prevention of cardiovascular diseases. J. Cardiovasc. Pharmacol. 2009, 54, 391−398. (79) Castañeda-Arriaga, R.; Vivier-Bunge, A.; Raul Alvarez-Idaboy, J. Primary antioxidant and metal-binding effects of tiopronin: A theoretical investigation of its action mechanism. Comput. Theor. Chem. 2016, 1077, 48−57. (80) Glazer, A. M.; Sofen, B. D.; Rigel, D. S.; Shupack, J. L. Successful treatment of generalized essential telangiectasia with 6mercaptopurine. J. Drugs Dermatol. 2017, 16, 280−282. (81) Chen, X. L.; She, S. F.; Li, Y. H.; Zhang, W. J.; Tang, Y. T.; Zhuang, K. H.; Pan, Y. B.; Liu, F. B.; He, W. L. Azathioprine/6mercaptopurine versus 5-aminosalicylic for treatment of inflammatory bowel disease. Int. J. Clin. Exp. Med. 2017, 10, 1927−1936. (82) Korelitz, B. I. Expert opinion: Experience with 6-mercaptopurine in the treatment of inflammatory bowel disease. World J. Gastroenterol. 2013, 19, 2979−2984. (83) Hübener, S.; Oo, Y. H.; Than, N. N.; Hübener, P.; WeilerNormann, C.; Lohse, A. W.; Schramm, C. Efficacy of 6Mercaptopurine as Second-Line Treatment for Patients With Autoimmune Hepatitis and Azathioprine Intolerance. Clin. Gastroenterol. Hepatol. 2016, 14, 445−453. (84) Li, G.-X.; Liu, Z.-Q. Captopril and 6-mercaptopurine: Whose SH possesses higher antioxidant ability? Eur. J. Med. Chem. 2009, 44, 4841−4847. (85) Legué, C.; Legros, L.; Kammerer-Jacquet, S.; Jézequel, C.; Houssel-Debry, P.; Uguen, T.; Le Lan, C.; Guillygomarc’h, A.;

Moirand, R.; Turlin, B.; Guyader, D.; Bardou-Jacquet, E. Safety and Efficacy of 6-thioguanine as a Second-line Treatment for Autoimmune Hepatitis. Clin. Gastroenterol. Hepatol. 2018, 16, 290−291. (86) de Boer, N. K. H.; van Nieuwkerk, C. M. J.; Aparicio Pages, M. N.; de Boer, S. Y.; Derijks, L. J. J.; Mulder, C. J. J. Promising treatment of autoimmune hepatitis with 6-thioguanine after adverse events on azathioprine. Eur. J. Gastroenterol. Hepatol. 2005, 17, 457−461. (87) Romagosa, R.; Kerdel, F.; Shah, N. Treatment of psoriasis with 6-thioguanine and hepatic venoocclusive disease. J. Am. Acad. Dermatol. 2002, 47, 0970−0972. (88) Sherer, D. W.; Lebwohl, M. G. 6-Thioguanine in the treatment of psoriasis: A case report and literature review. J. Cutaneous Med. Surg. 2002, 6, 546−550. (89) Zackheim, H. S.; Glogau, R. G.; Fisher, D. A.; Maibach, H. I. 6Thioguanine treatment of psoriasis: Experience in 81 patients. J. Am. Acad. Dermatol. 1994, 30, 452−458. (90) Chaabane, W.; Appell, M. L. Interconnections between apoptotic and autophagic pathways during thiopurine-induced toxicity in cancer cells: The role of reactive oxygen species. Oncotarget 2016, 7, 75616−75634. (91) Cuin, A.; Massabni, A. C.; Pereira, G. A.; Leite, C. Q. F.; Pavan, F. R.; Sesti-Costa, R.; Heinrich, T. A.; Costa-Neto, C. M. 6Mercaptopurine complexes with silver and gold ions: Anti-tuberculosis and anti-cancer activities. Biomed. Pharmacother. 2011, 65, 334−338. (92) Issaeva, N.; Thomas, H. D.; Djurenovic, T.; Jaspers, J. E.; Stoimenov, I.; Kyle, S.; Pedley, N.; Gottipati, P.; Zur, R.; Sleeth, K.; Chatzakos, V.; Mulligan, E. A.; Lundin, C.; Gubanova, E.; Kersbergen, A.; Harris, A. L.; Sharma, R. A.; Rottenberg, S.; Curtin, N. J.; Helleday, T. 6-thioguanine selectively kills BRCA2-defective tumors and overcomes PARP inhibitor resistance. Cancer Res. 2010, 70, 6268−6276. (93) Cara, C. J.; Pena, A. S.; Sans, M.; Rodrigo, L.; Guerrero-Esteo, M.; Hinojosa, J.; García-Paredes, J.; Guijarro, L. G. Reviewing the mechanism of action of thiopurine drugs: Towards a new paradigm in clinical practice. Med. Sci. Monit. 2004, 10, RA247−RA254. (94) Misdaq, M.; Ziegler, S.; Von Ahsen, N.; Oellerich, M.; Asif, A. R. Thiopurines induce oxidative stress in T-lymphocytes: A proteomic approach. Mediat. Inflamm. 2015, 2015, 434825. (95) Pelin, M.; De Iudicibus, S.; Londero, M.; Spizzo, R.; Dei Rossi, S.; Martelossi, S.; Ventura, A.; Decorti, G.; Stocco, G. Thiopurine biotransformation and pharmacological effects: Contribution of oxidative stress. Curr. Drug Metab. 2016, 17, 542−549. (96) Flora, S. J. S.; Pachauri, V. Chelation in metal intoxication. Int. J. Env. Res. Public Health 2010, 7, 2745−2788. (97) Kalia, K.; Flora, S. J. S. Strategies for safe and effective therapeutic measures for chronic arsenic and lead poisoning. J. Occup. Health 2005, 47, 1−21. (98) Pandey, S.; Sharma, V.; Chaudhary, A. K. Chelation therapy and chelating agents of Ayurveda. Int. J. Green Pharm. 2016, 10, 143− 150. (99) Tian, R.; Shi, R. Dimercaprol is an acrolein scavenger that mitigates acrolein-mediated PC-12 cells toxicity and reduces acrolein in rat following spinal cord injury. J. Neurochem. 2017, 141, 708−720. (100) Derk, C. T.; Huaman, G.; Jimenez, S. A. A retrospective randomly selected cohort study of D-penicillamine treatment in rapidly progressive diffuse cutaneous systemic sclerosis of recent onset. Br. J. Dermatol. 2008, 158, 1063−1068. (101) Khanna, D.; Furst, D. E.; Hays, R. D.; Park, G. S.; Wong, W. K.; Seibold, J. R.; Mayes, M. D.; White, B.; Wigley, F. F.; Weisman, M.; Barr, W.; Moreland, L.; Medsger, T. A., Jr; Steen, V. D.; Martin, R. W.; Collier, D.; Weinstein, A.; Lally, E. V.; Varga, J.; Weiner, S. R.; Andrews, B.; Abeles, M.; Clements, P. J. Minimally important difference in diffuse systemic sclerosis: Results from the Dpenicillamine study. Ann. Rheum. Dis. 2006, 65, 1325−1329. (102) Członkowska, A.; Litwin, T.; Karliński, M.; Dziezyc, K.; Chabik, G.; Czerska, M. D-penicillamine versus zinc sulfate as firstline therapy for Wilson’s disease. Eur. J. Neurol. 2014, 21, 599−606. 18590

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591

ACS Omega

Article

(103) Delangle, P.; Mintz, E. Chelation therapy in Wilson’s disease: From d-penicillamine to the design of selective bioinspired intracellular Cu(i) chelators. Dalton Trans. 2012, 41, 6359−6370. (104) Lowette, K. F.; Desmet, K.; Witters, P.; Laleman, W.; Verslype, C.; Nevens, F.; Fevery, J.; Cassiman, D. M. Wilsons disease: long-term follow-up of a cohort of 24 patients treated with Dpenicillamine. Eur. J. Gastroenterol. Hepatol. 2010, 22, 564−571. (105) Capitani, E. M. Diagnosis and treatment of lead poisoning in children and adults. Medicina 2009, 42, 319. (106) Havre, P. A.; O’Reilly, S.; McCormick, J. J.; Brash, D. E. Transformed and tumor-derived human cells exhibit preferential sensitivity to the thiol antioxidants, N-acetyl cysteine and penicillamine. Cancer Res. 2002, 62, 1443−1449. (107) Martínez, A.; Vargas, R.; Galano, A. How to identify promising metal scavengers? d-penicillamine with copper as a study case. Int. J. Quantum Chem. 2018, 118, No. e25457. (108) Qiao, S.; Cabello, C. M.; Lamore, S. D.; Lesson, J. L.; Wondrak, G. T. D-Penicillamine targets metastatic melanoma cells with induction of the unfolded protein response (UPR) and Noxa (PMAIP1)-dependent mitochondrial apoptosis. Apoptosis 2012, 17, 1079−1094. (109) Squitti, R.; Rossini, P. M.; Cassetta, E.; Moffa, F.; Pasqualetti, P.; Cortesi, M.; Colloca, A.; Rossi, L.; Finazzi-Agro’, A. Dpenicillamine reduces serum oxidative stress in Alzheimer’s disease patients. Eur. J. Clin. Invest. 2002, 32, 51−59. (110) Jalali, S. M.; Najafzadeh, H.; Bahmei, S. Protective role of silymarin and D-penicillamine against lead-induced liver toxicity and oxidative stress. Toxicol. Ind. Health 2017, 33, 512−518. (111) Smith, R. M.; Martell, A. E.; Motekaitis, R. J. NIST Standard Reference Database 46. NIST Standard Reference Database 46, 8.0 ed.; National Institute of Standards and Technology (NIST): Texas A&M University, 2004. (112) Freisleben, H.-J.; Tandian, D.; Kuebel-Thiel, K. Application of α-lipoic acid beyond peripheral diabetic polyneuropathy (A report of 3 cases). Med. J. Indones. 2011, 20, 143−148. (113) Chen, W.; Shockcor, J. P.; Tonge, R.; Hunter, A.; Gartner, C.; Nelson, S. D. Protein and Nonprotein Cysteinyl Thiol Modification byN-Acetyl-p-benzoquinone Imine via a NovelIpsoAdduct. Biochemistry 1999, 38, 8159−8166. (114) Galano, A.; Medina, M. E.; Tan, D. X.; Reiter, R. J. Melatonin and its metabolites as copper chelating agents and their role in inhibiting oxidative stress: a physicochemical analysis. J. Pineal Res. 2014, 58, 107−116. (115) Á lvarez-Diduk, R.; Galano, A. Adrenaline and noradrenaline: Protectors against oxidative stress or molecular targets? J. Phys. Chem. B 2015, 119, 3479−3491. (116) Lipinski, C. A.; Lombardo, F.; Dominy, B. W.; Feeney, P. J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings 1PII of original article: S0169-409X(96)00423-1. The article was originally published in Advanced Drug Delivery Reviews 23 (1997) 3-25. 1. Adv. Drug Delivery Rev. 2001, 46, 3−26. (117) Ghose, A. K.; Viswanadhan, V. N.; Wendoloski, J. J. A knowledge-based approach in designing combinatorial or medicinal chemistry libraries for drug discovery. 1. A qualitative and quantitative characterization of known drug databases. J. Comb. Chem. 1999, 1, 55−68. (118) Veber, D. F.; Johnson, S. R.; Cheng, H.-Y.; Smith, B. R.; Ward, K. W.; Kopple, K. D. Molecular properties that influence the oral bioavailability of drug candidates. J. Med. Chem. 2002, 45, 2615− 2623. (119) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi,

R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. (120) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Design of density functionals by combining the method of constraint satisfaction with parametrization for thermochemistry, thermochemical kinetics, and noncovalent interactions. J. Chem. Theory Comput. 2006, 2, 364−382. (121) Galano, A.; Alvarez-Idaboy, J. R. Kinetics of radical-molecule reactions in aqueous solution: A benchmark study of the performance of density functional methods. J. Comput. Chem. 2014, 35, 2019− 2026. (122) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378−6396. (123) Eyring, H. The activated complex in chemical reactions. J. Chem. Phys. 1935, 3, 107−115. (124) Evans, M. G.; Polanyi, M. Some applications of the transition state method to the calculation of reaction velocities, especially in solution. Trans. Faraday Soc. 1935, 31, 875−894. (125) Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J. Current status of transition-state theory. J. Phys. Chem. 1996, 100, 12771− 12800. (126) Kuppermann, A.; Truhlar, D. G. Exact tunneling calculations. J. Am. Chem. Soc. 1971, 93, 1840−1851. (127) Garrett, B. C.; Truhlar, D. G. Semiclassical tunneling calculations. J. Phys. Chem. 1979, 83, 2921−2926. (128) Galano, A.; Alvarez-Idaboy, J. R. A computational methodology for accurate predictions of rate constants in solution: Application to the assessment of primary antioxidant activity. J. Comput. Chem. 2013, 34, 2430−2445. (129) Collins, F. C.; Kimball, G. E. Diffusion-controlled reaction rates. J. Colloid Sci. 1949, 4, 425−437. (130) Smoluchowski, M. Mathematical theory of the kinetics of the coagulation of colloidal solutions. Z. Physiol. Chem. 1917, 92, 129− 168. (131) Einstein, A. Ü ber die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten Teilchen. Ann. Phys. 1905, 322, 549−560. (132) Stokes, G. G. Mathematical and Physical Papers; Cambridge University Press: Cambridge, 1903; Vol. 3. (133) Pérez-González, A.; Castañeda-Arriaga, R.; Verastegui, B.; Carreón-González, M.; Alvarez-Idaboy, J. R.; Galano, A. Estimation of empirically fitted parameters for calculating pK a values of thiols in a fast and reliable way. Theor. Chem. Acc. 2017, 137, 5. (134) Galano, A.; Pérez-González, A.; Castañ eda-Arriaga, R.; Muñ oz-Rugeles, L.; Mendoza-Sarmiento, G.; Romero-Silva, A.; Ibarra-Escutia, A.; Rebollar-Zepeda, A. M.; León-Carmona, J. R.; Hernández-Olivares, M. A.; Alvarez-Idaboy, J. R. Empirically Fitted Parameters for Calculating pKa Values with Small Deviations from Experiments Using a Simple Computational Strategy. J. Chem. Inf. Model. 2016, 56, 1714−1724.

18591

DOI: 10.1021/acsomega.8b02943 ACS Omega 2018, 3, 18582−18591