Revealing the Mechanistic Pathway of Acid Activation of Proton Pump

Dec 5, 2016 - ... functions, including protein digestion, calcium and iron absorption, as well as providing some protection against bacterial infectio...
0 downloads 0 Views 1022KB Size
Subscriber access provided by Warwick University Library

Article

Revealing the Mechanistic Pathway of Acid Activation of Proton Pump Inhibitors to Inhibit the Gastric Proton Pump: A DFT Study Kalyanashis Jana, Tusar Bandyopadhyay, and Bishwajit Ganguly J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.6b09334 • Publication Date (Web): 05 Dec 2016 Downloaded from http://pubs.acs.org on December 12, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry B is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Revealing the Mechanistic Pathway of Acid Activation of Proton Pump Inhibitors to Inhibit the Gastric Proton Pump: A DFT Study Kalyanashis Jana1, Tusar Bandyopadhyay2 and Bishwajit Ganguly1* 1

Computation and Simulation Unit (Analytical Discipline and Centralized Instrument Facility), CSIR–Central

Salt and Marine Chemicals Research Institute, Bhavnagar–364002, Gujarat, India and Academy of Scientific and Innovative Research, CSIR–CSMCRI, Bhavnagar–364002, Gujarat, India. 2

Theorectical Chemistry Section, Chemistry Group MOD LAB, Bhabha Atomic Research Centre, Trombay,

Mumbai 400 085, INDIA * To whom correspondence should be addressed. E-mail: [email protected]

Abstract: Acid-related gastric diseases are associated with disorder of digestive tract acidification due to the acid secretion by gastric proton pump, H+,K+-ATPase. Omeprazole is one of the persuasive irreversible inhibitor of the proton pump H+,K+-ATPase. However, the reports on the mechanistic pathway of irreversible proton pump inhibitors (PPIs) on the acid activation and formation of disulfide complex are scarce in the literature. We have examined the acid activation PPIs i.e. timoprazole, S-omeprazole and R-omeprazole using M062X/631++G(d,p) in aqueous phase with SMD solvation model. The proton pump inhibitor is a prodrug and activated in the acidic canaliculi of the gastric pump H+,K+-ATPase to sulfenic acid which can either form another acid activate intermediate sulfenamide or a disulfide complex with cysteine amino acid of H+,K+-ATPase. The quantum chemical calculations suggest that the transition state (TS5) for the disulfide complex formation is the rate determining step of the multistep acid inhibition process by PPIs. The free energy barrier of TS5 is 5.5 kcal/mol higher for timoprazole compared to the S-omeprazole. The stability of the transition state for the formation of disulphide bond between S-omeprazole and cysteine

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

amino acid of H+,K+-ATPase is governed by inter- and intramolecular hydrogen bonding. The disulphide complex for S-omeprazole is thermodynamically more stable by 4.5 kcal/mol in aqueous phase compared to disulfide complex of timoprazole, which corroborates the less efficacy of timoprazole as irreversible PPI for acid inhibition process. It has been speculated that sulfenic acid can either form sulfenamide or a stable disulfide complex with cysteine amino acid residue of H+,K+-ATPase. The M062X/6-31++G(d,p) level of theory calculated results reveal that the formation of tetra cyclic sulfenamide is unfavoured by ~ 17 kcal/mol for S-omeprazole and 11.5 kcal/mol for timoprazole compared to the disulfide complex formation in each case. The DFT calculations have further shed light on the acid activation process of R- and S-isomers of omeprazole. The calculated results suggest that the efficacy of these isomers lie on their metabolic pathway and excretion from human body.

Introduction: In 1905, John Edkins discovered gastrin that has subsequently stimulated the scientific exploration of the secretion of gastric acid.1 Gastric acid plays a key role in normal upper gastrointestinal functions, including protein digestion, calcium and iron absorption, as well as providing some protection against bacterial infections.2 However, inappropriate levels of gastric acid in stomach and intestine cause several pathological conditions, e.g. Gastroesophageal Reflux Disease (GERD) and peptic ulcer (PU).2-6 Heartburn is the one of the most common symptom for GERD and peptic ulcer causes pain in the stomach and small intestine.4-6 Discovery of gastrin and elucidation of the role of histamine led to the understanding of the pathogenic basis of PU, GERD and gastric acid related diseases and their subsequent treatments. The secretion gastric acid in parietal cells is due to the presence of food in stomach or intestine and also due to the taste, smell, sight or thought of food.2,7-8 Gastric acid is secreted followed by the activation of histamine, acetylcholine or gastrin receptors (the H2, M3 and 2 ACS Paragon Plus Environment

Page 2 of 24

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CCK2 receptors, respectively) located in the basolateral membrane of the parietal cells. The H2, M3 and CCK2 receptors initiate signal transduction pathways that triggers the activation of the H+,K+-ATPase which causes acid secretion.8 Acid-related gastric diseases are associated with disorder of digestive tract acidification. The gastric proton pump, H+,K+ATPase, exports H+ in the luminal cavity in exchange of K+, which generates a highly acidic environment in the gastrointestinal track, and therefore, is a target for acid suppressant drug molecules. This perhaps the commonly used pharmacological approach to regulate the acid secretion as it can inhibit acid secretion independently. Omeprazole is known to be the first proton-pump inhibitors used in clinical practice.2,9 The development of omeprazole was a new approach for the effective inhibition of acid secretion and the treatment of acid-related diseases, and was quite quickly shown to be clinically superior to the H2-receptor antagonists. Omeprazole consists of two optical isomers (enantiomers), one being the mirror image of the other, and exists as racemic mixture.11-12 The S-isomer of omeprazole subsequently proved to be a better drug that is significantly superior compared to the racemic omeprazole. Pharmacokinetics and pharmacodynamics studies have revealed that the Sisomer of the proton pump inhibitors are more superior compared to the R-isomers.12-13 Like omeprazole drug, other irreversible proton pump inhibitors e.g. timoprazole, tenatoprazole, rabeprazole, pantoprazole, lansoprazole are racemates in nature.14 It is important to note that omeprazole and other proton-pump inhibitors are prodrugs which are converted to their active form in acidic environments of parietal cell.15-21 Proton pump inhibitors are weak bases and generally concentrate in the acidic secretory canaliculi of the parietal cell, where they are activated to an acid activated complex by a proton-catalysed process (Scheme 115).22 The acid activated complex makes covalent bond with the sulphydryl groups of cysteine-813 or 822 residues in the extracellular domain of the H+,K+-ATPase to generate stable disulfide complex and thereby inhibiting the enzyme activity.

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme1: Mechanism of the acid activation and finally stable disulfide complex formation of proton pump inhibitor

Scheme 1 shows the plausible mechanism to inhibit the acid secretion by PPI.15 The Scheme 1 suggests that pyridine (Py) nitrogen of neutral drug molecule gets protonated in the acidic environment in the parietal cell to form (PyH+) intermediate. Further, the N-3 nitrogen of benzimidazole (Bz) moiety is protonated via the transfer of proton from the pyridinium nitrogen (Scheme 1 and 2). The nitrogen lone-pair of pyridine unit intramolecularly attacks the C-2 carbon of the Bz moiety to form the N-C bond and a spiro intermediate i.e., dihydrobenzimidazole is formed. The unstable spiro intermediate opens up and undergoes aromatization to form sulfenic acid which further dehydrates to form tetracyclic sulfenamide (Scheme 1). Finally, the sulfenic acid or the tetracyclic sulfenamide reacts with the cysteine

4 ACS Paragon Plus Environment

Page 4 of 24

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

residue of H+,K+-ATPase enzyme to form the stable disulfide bond that suppress the acid secretion process. The plausible mechanism for the inhibition of acid secretion with PPI has not been examined experimentally or computationally. In this article, we have explored the mechanism of disulfide bond formation between the cysteine amino acid residue of the gastric proton pump H+,K+-ATPase and PPIs. Timoprazole one of the PPIs is the basic scaffold of many irreversible PPIs.23,15 The derivatization of timoprazole shows better efficacy compared to parent temiprazole.2,24-25 Therefore, we have examined the acid activation mechanism with timoprazole and one of the superior PPI, i.e. omeprazole.2 5-methoxy-2-(((4-methoxy-3,5dimethylpyridin-2-yl)methyl)sulfinyl)-1H-benzo[d]imidazole (omeprazole) is a substituted 2-((pyridin-2-ylmethyl)sulfinyl)-1H-benzo[d]imidazole (timoprazole) (Scheme 2). It would be interesting to see that such structural difference in PPIs can induce the activity toward the suppression of acid secretion in the acidic canaliculi. Further, the experimental observations fail to shed light on the controversy of disulfide bond formation with sulfenic acid or with sulfenamide of the drug molecules. The computational study showed a clear preference for the formation of disulphide bond between the enzyme and sulfenic acid. This is the first report to decipher this ambiguous observation in the experimental studies of PPI.

Scheme 2: a) S-isomer of the timoprazole drug molecule b) S-isomer of the omeprazole drug i.e. Someprazole and c) R-isomer of omeprazole.

Computational Methods: 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

All the acid activated geometries of timoprazole, R- and S-omeprazole have been optimised in gas phase at M062X26 DFT functional with 6-31++G(d,p)27-28 Pople basis set. We have performed harmonic frequency calculations at the same level of theory to confirm minima of optimised geometries with no imaginary frequencies. The M062X is one of the accurate DFT functional for organic and ionic systems.29-30 We have located transition states on the potential energy surface with M062X/6-31++G(d,p) in gas phase and a transition state is confirmed with one imaginary frequency. Further, we have performed IRC calculation to connect the TS geometries with the initial and final complex.31 The electronic energy differences were calculated with respect to the initial neutral drug molecule as:

ΔE =EX– EN ........ (1) where EX is electronic energy of intermediates, transition states or final complex and EN is the electronic energy of neutral drug molecule molecules and ΔE is the difference in the electronic energies. The M062X/6-31++G(d,p) level of theory optimized geometries were taken for single point energy calculations in aqueous phase with Self Consistent Reaction Field (SCRF) method.32 We have performed the aqueous phase (ε=78.8) calculation using SMD solvation model with M062X/6-31++G(d,p) level of theory.33 It has been reported that SMD solvation model is a universal solvation model, where “universal” denotes its applicability to any charged or uncharged solute in any solvent or liquid medium for which a few key descriptors are known. We have calculated the free energy differences using similar equation for acid activation process of the PPIs. The Gibbs free energy in the solvent phase has been calculated as: ‫ܩ‬௔௤ = ‫ܩ‬௚௔௦ – (‫ܧ‬௔௤ – ‫ܧ‬௚௔௦ ) + ߂‫ ܩ‬ଵ௔௧௠→ଵெ …….2 where Gaq and Eaq are Gibbs free energy and electronic energy in the aqueous phase, Ggas and Egas are Gibbs free energy and electronic energy in the gas phase, ߂‫ ܩ‬ଵ௔௧௠→ଵெ is the the

6 ACS Paragon Plus Environment

Page 6 of 24

Page 7 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

correction associated with the change in the standard state from the gas phase (1 atm) to solution (1 mol L−1) and its value at 298.15 K is 0.003012 Hatree.34-36 Free energy differences have been calculated as: ߂‫ܩ‬௔௤ = ‫ܩ‬௔௤,௑ − ‫ܩ‬௔௤,ே − ‫ܩ‬௔௤,ு శ ...............3 where Gaq,X is the free energy of the intermediates, transition states, Gaq,N is the free energy of the neutral drug molecules and ‫ܩ‬௔௤,ு శ is the free energy of the proton. The ‫ܩ‬௔௤,ு శ has been calculated as:37 ଴ ଵ௔௧௠→ଵெ ‫ܩ‬௔௤,ு శ = ‫ܩ‬௚௔௦,ு ...............4 శ + ߂‫ܩ‬௔௤,௦௢௟,ு శ + ߂‫ܩ‬

38 ଴ where ‫ܩ‬௚௔௦,ு kcal/mol and ߂‫ ܩ‬ଵ௔௧௠→ଵெ =1.89 శ = − 6.28 kcal/mol, ߂‫ܩ‬௔௤,௦௢௟,ு శ = − 261.85

kcal/mol are obtained from literature. The ߂‫ܩ‬௔௤,௦௢௟,ு శ is debated in the literature and difficult to either measure or predict accurately. In

this article, we have considered

߂‫ܩ‬௔௤,௦௢௟,ு శ =−261.85 kcal/mol and the ‫ܩ‬௔௤,ு శ =−266.24 kcal/mol value has been used in (3).38 Further, we have additionally performed aqueous phase calculations of all the acid activated intermediates and transition states of timoprazole drug molecule using the same level of theory. The free energy differences calculated using zero-point energy as well as the entropy term for timoprazole using Equation-3 (Table S1). We have observed that ΔGaq obtained using Gaq calculated by Equation-2 and the optimization of timoprazole in aqueous phase showed similar results. All DFT calculations were performed using the G09 package. 39

Results and Discussion: The mechanism of acid activation and disulphide bond formation with cysteine of H+K+ ATPASE has been examined with M062X/6-31++G(d,p) level of theory in aqueous 7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 24

phase. To investigate the effect of substituents on the overall acid activation process timoprazole and S-omeprazole molecules have been considered. The basic scaffold of timoprazole has been substituted with the methoxy groups at the 4-position in pyridine (Py) and 5-position benzimidazole (Bz) rings and with methyl group at 3, 5 position of the Py in Someprazole molecule (Scheme 2). The M062X/6-31++G(d,p) calculated results show that the pyridine Py and Bz rings are co-planner for timoprazole molecule (Figure 2). Such arrangement of rings facilitates the proton transfer process in the inhibitor molecule to form the acid activated intermediate (Figure 2). The oxygen of the sulfinyl group of timoprazole forms a weak intramolecular hydrogen bond (2.37 Ǻ) with N-H proton of the Bz unit and the bond angle between the -O--H-N is ~103º. The protonation free energy of pyridine nitrogen is -3.8 kcal/mol compared to the neutral timoprazole drug molecule (Figure 1 and Table 1). The protonated intermediate (Timo-PyH+) also experiences the intramolecular hydrogen bonding interaction between the oxygen of sulfinyl group and N-H proton of the benzimidazole moiety. The transition state free energy barrier (Timo-TS1) for the protonation of the benzimidazole nitrogen has been found to be 4.2 kcal/mol compared to preceding TimoPyH+ and the intermediate Timo-BzH+ is less stable by 0.4 kcal/mol compared to the neutral drug molecule. The free energy profile suggests that the formation of Timo-BzH+ could be barrierless, where Timo-BzH+ and Timo-TS1 are energetically higher by 0.4 kcal/mol than timoprazole, however, electronic energy (∆Eaq) results show that Timo-BzH+ is stable by 1.4 kcal/mol compared to Timo-TS1 (Table S2). On the other hand, there are

reports

available in the literature, which suggests that the protonation of the Bz moiety might be favorable compared to the Py unit.16,40 Therefore, we have examined the pKa of both Bz and Py rings using the M062X/6-31++G(d,p) level of theory. The calculated pKa values suggest that the N1 nitrogen of the Py moiety is more basic than the N3 nitrogen of the Bz unit which is in agreement with the experimental report of Shin et al. (Table S3).15

8 ACS Paragon Plus Environment

Page 9 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1: Free energy profile of the acid activation process further disulfide complex formation of the timoprazole drug molecule.

Further, pyridine nitrogen attacks the C-2 carbon atom of the Bz moiety (Scheme 2), which leads to the formation of the transition state (Timo-TS2) with free energy barrier of 13.9 kcal/mol compared to the preceding intermediate Timo-BzH+ and consequently yields spiro intermediate dihydrobenzimidazole (Timo-spiro, Figure 2). The DFT calculated distance between C-2 carbon of Bz moiety and nitrogen of Py unit in the transition state (Timo-TS2, Figure 2) geometry is 1.90 Ǻ. The spiro complex is also an unstable intermediate compared to the timoprazole drug molecule (5.7 kcal/mol). Table 1: The M062X/6-31++G(d,p) level of theory calculated free energy difference calculated using Equation-3 given in kcal/mol. PyH+ Timoprazole S-Omeprazole R-Omeprazole

∆Gaq ∆Gaq ∆Gaq

-3.8 -5.4 -3.6

TS1 0.4 -0.5 -0.2

BzH+ 0.4 -1.9 -1.0

TS2 14.3 11.1 10.0

Spiro 5.7 1.7 2.6

TS3 11.4 7.1 7.6

Sulfenic Acid -9.9 -12.6 -10.1

9 ACS Paragon Plus Environment

TS4 51.6 48.8 50.2

Sulfenamide -14.9 -17.3 -17.9

TS5 40.1 31.9 32.4

Disulfide -29.4 -33.9 -30.4

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The N-H proton of Bz ring transferred to the sulfinyl oxygen to form sulfenic acid and the computed free energy barrier is 5.7 kcal/mol compared to the previous spiro intermediate (Figure 2). The corresponding intermediate sulfenic acid is 9.9 kcal/mol lower compared the neutral drug molecule. Intramolecular hydrogen bonding interaction N-H---O (1.93 Ǻ) was observed, which presumably stabilizes the sulfenic acid intermediate. Dehydration of sulfenic acid can lead to the formation of tetracyclic sulfenamide, however, activation barrier for such dehydration process is 61.5 kcal/mol compared to the sulfenic acid intermediate. The DFT calculations suggest that the tetra cyclic sulfenamide is stable by only 5.0 kcal/mol compared to the sulfenic acid. The water elimination and S-N bond formation occurs in a concerted way through a four membered ring formation in the transition state geometry (Timo-TS4, Figure 2). On the other hand, the acid activated complex sulfenic acid can form stable disulfide bond with cysteine amino acid of H+,K+-ATPase followed by the elimination of water molecule, which is the desired complex for the acid inhibition process (Figure 2). We have considered neutral cysteine amino acid to mimic the cysteine amino residue of the gastric proton pump H+,K+ATPase, whereas zwitterionic form of cysteine might exist, however, as discrete amino acid.41 The free energy barrier for the disulphide complex formation (Timo-TS5) is 50.0 kcal/mol compared to the sulfenic acid and disulfide complex is remarkably stable (-29.4 kcal/mol) compared to the tetra cyclic sulfenamide (Figure 1). Dehydration and S-S bond formation takes place through the formation of four membered transition state (Timo-TS5) and intramolecular hydrogen bond N-H---O (2.00 Ǻ) becomes weaker compared to timoprazole sulfenic acid. The DFT calculated results reveal that the formation of disulfide complex with cysteine amino acid of H+,K+-ATPase is kinetically and thermodynamically more stable than sulfenamide intermediate.

10 ACS Paragon Plus Environment

Page 10 of 24

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2: Mechanism of the acid activation and finally stable disulfide complex formation of timoprazole proton pump inhibitor.

We have extended our quantum chemical calculations with the potent proton pump inhibitor S-omeprazole drug molecule. We have performed the calculation in aqueous phase with M062X/6-31++G(d,p) level of theory. Protonation of pyridine nitrogen of S-omeprazole forms S-Ome-PyH+ and the protonation free energy of S-omeprazole is -5.4 kcal/mol (Figure 3 and 4). The PES of S-omeprazole showed the similar pathway to form sulfenic acid as observed with timoprazole (Figure 2 and 4).

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: Free energy profile of the acid activation process and further disulfide complex formation of the S-omeprazole drug molecule.

S-Ome-sulfenic acid is stable by -12.6 kcal/mol compared to the S-omeprazole in the free energy surface in the acid activation process. Further, dehydration can either form a tetracyclic sulfenamide intermediate or a stable disulfide complex with cysteine amino acid of H+,K+-ATPase through a four membered ring transition state (S-Ome-TS4/5). The free energy barrier (S-Ome-TS4) for the formation of sulfenamide is 61.4 kcal/mol compared to the preceding intermediate sulfenic acid and dehydrated sulfenamide is stable by only -4.7 kcal/mol compared to the sulfenic acid.

12 ACS Paragon Plus Environment

Page 12 of 24

Page 13 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4: Acid activation and further disulfide bond formation mechanistic pathway of the Someprazole drug molecule.

On the other hand, the disulfide bond formation with cysteine amino acid of H+,K+ATPase by dehydration is kinetically and thermodynamically more favorable compared to the sulfenamide formation (Figure 3). The free energy barrier calculated for the disulfide bond formation is 44.5 kcal/mol compared to the previous intermediate S-Ome-sulfenic acid for Someprazole molecule.

The transition state (TS5) is the rate determining step for the

formation of disulphide complex between the PPI and the cysteine residue of H+,K+-ATPase in both timoprazole and S-omeprazole systems. It is to note that the activation barrier for the formation of (TS5) is lower (5.5 kcal/mol) in the case of S-omeprazole compared to 13 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

timoprazole (Figure 1, 2, 3 and 4). The stability of TS5 S-omeprazole is governed by the hydrogen bonding interactions between the PPI and cysteine molecule. The intermolecular hydrogen bonding interaction (N-H---O, 1.8 Ǻ) between carboxylic oxygen of the cysteine amino acid residue and N-H proton of Bz moiety was observed and the intramolecular Hbonding interaction (C-H---N, 2.29 Ǻ) was also observed between the methylene proton of sulfenic acid and N-1 nitrogen of Bz unit. The torsional angle calculated (N3-C2-N1’-C2’) for S-omeprazole is ~48º, which is much smaller compared to the corresponding torsional angle of ~60º (Figure S2). Analysing the geometry of sulfenic acid and the TS5 for timoprazole and omeprazole suggests that the substituents at remote position of the Bz and Py moiety helps to align the two aromatic rings (Py and Bz) in S-omeprazole. Such a difference in the rate determining activation step can augment the efficacy of S-omeprazole compared to timoprazole molecule in acid inhibition process. The disulfide complex formed in the acidic environment of the parietal cell is energetically preferred compared to the corresponding tetra cyclic sulfenamide intermediate. The disulfide complex formed from the S-omeprazole is stable by -33.9 kcal/mol compared to the S-omeprazole drug molecule. The formation of the tetracyclic sulfenamide is less likely in this case presumably due to the formation of puckered six membered ring through a four membered transition state geometry (S-Ome-TS4). It has been reported that S-isomer i.e. S-omeprazole is more efficient compared to the R-isomer of omeprazole.42,12-14 The pharmacokinetic and pharmacodynamics suggests that the activity of R- and S-isomer of omeprazole might differ in the metabolic pathway.43-44 However, the mechanistic pathway for the formation of disulfide complex between cysteine and R- and S-isomer of omeprazole remain unexplored. At the cellular level, both R- and Sisomers are protonated and converted in the acidic compartment of the parietal cell in exactly the same way to form acid activated achiral sulfenic acid or tretra cyclic sulphonamide which 14 ACS Paragon Plus Environment

Page 14 of 24

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

makes disulfide bond with thiol group of the Cys-813 to inhibit H+,K+-ATPase.42 We have computed the acid activation and the disulfide bond formation process of omeprazole at molecular level using M062X/6-31++G(d,p) level of theory in aqueous phase. We have not also observed any significant difference in the acid activation and disulfide complex formation (Table 1, Scheme S1, Figure S1 and Figure S3). The free energy differences with respect to neutral omeprazole show that the free energy difference for the acid activation process and final disulfide complex formation is similar for these two isomers i.e. the acid activation of the R-omeprazole and S-omeprazole is equally favorable in the acidic canaliculi of the parietal cell (Table 1). Therefore, it appears that the acid activation process of R- and S- isomers of omeprazole starts to differ in their metabolic pathway and excretion from human body.

Conclusion: Irreversible acid suppressants are common therapeutic option for the treatment of acid-related diseases since the discovery of omeprazole. However, the reports on the mechanism of acid activation and the inhibition of these prodrugs are scarce in literature. In this work, we have reported the mechanistic pathway for the acid activation and the disulfide bond formation with PPIs i.e. timoprazole, S-omeprazole and R-omeprazole computationally with M062X/6-31++G(d,p) level of theory. The effect of solvation has been studied with SCRF method using single point SMD solvation model in aqueous phase. The reactant, intermediates, transition states, and products on the potential energy surface of timoprazole was also optimized using the same solvation model and the results were found to be similar in both cases. The acid activation and the inhibition of omeprazole is superior over timoprazole molecule. Omeprazole is the substituted derivative of timoprazole and hence warranted a study to shed light on the efficacy of these drug molecules. The DFT calculations reveal that

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the acid activation of the PPIs and further gastric proton pump inhibition process follow a multistep process in which the drug molecules form a disulphide bond with the cysteine residue of H+,K+-ATPase. The free energy surface calculated for timoprazole suggests that the disulphide bond formation is energetically unfavored compared to omeprazole. The transition state (TS5) for the formation of disulfide complex is the rate determining step and one of the key steps of the acid inhibition process by PPIs. The TS5 free energy barrier for timoprazole is 5.5 kcal/mol higher than S-omeprazole. The stabilization of TS5 for Someprazole occurs due to strong inter- and intramolecular hydrogen bonding while forming the disulphide bond with cysteine residue. Such interactions and the remote substituents further helps to align the two aromatic rings (Py and Bz) in S-omeprazole to achieve more planarity compared to timoprazole molecule and hence to augment the stability. Further, the disulphide complex for S-omeprazole is also stable by 4.5 kcal/mol in aqueous phase compared to disulfide complex of timoprazole. These computed results corroborate the less efficacy of timoprazole compared to S-omeprazole for irreversible acid suppression process. The experimental observations indicate that during the acid activation process sulfenic acid can either produce sulfenamide or a stable disulfide complex with cysteine of H+,K+ATPase. The DFT results clearly suggest that the free energy barrier for the formation of tetra cyclic sulfenamide is energetically unfavoured by ~ 17 kcal/mol for S-omeprazole and 11.5 kcal/mol for timoprazole compared to the disulfide complex formation in each case. The reaction energies also show that the desired disulfide complex is ~15 kcal/mol more stable than the acid activated intermediate sulfenamide. The M062X/6-31++G(d,p) level of theory calculated results further suggests that the acid activation process of R- and S-isomers of omeprazole and their inhibition effect is very similar and the efficacy of S-omeprazole presumably depends on metabolic pathway and

16 ACS Paragon Plus Environment

Page 16 of 24

Page 17 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

excretion from human body. Further work is under progress to examine the metabolic pathways of R- and S-isomers of omeprazole.

ASSOCIATED CONTENT: Supporting Information M062X/6-31++G(d,p) level of theory optimized Cartesian coordinates have been given in SI. The M062X/6-31++G(d,p) theory calculated aqueous phase and gas electronic energy and calculated pKa values have been given in SI. The acid activation mechanism and free energy profile of R-Omeprazole is also available in SI. This information is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION: Corresponding Author *Fax:

(+91)-278-2567562.

Telephone:

+91-278-2567760,

ext

6770.

E-mail:

[email protected]

ACKNOWLEDGMENTS: CSMCRI communication no: 089/ 2016. K. J. is thankful to UGC, New Delhi, India, for awarding a senior research fellowship. K. J. acknowledges to AcSIR for his PhD registration. B.G. thanks MSM, SIP (CSIR, New Delhi) and BRNS, Mumbai, for financial support. This work is supported by Department of Atomic Energy, Government of India, Board

of

Research

in

Nuclear

Sciences

(DAE-BRNS)

under

grant

no.

2013/37C/54/BRNS/2278. K. J. would like to thank Dr. Rabindranath Lo for helpful discussions and valuable suggestions to design this project. We thank the anonymous reviewer’s for their valuable comments/suggestions that have helped us to improve the paper.

References:

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1) Modlin, I. M.; Sachs, G.; Wright, N.; Kidd, M. Edkins and a Century of Acid Suppression. Digestion 2005, 72, 129–145. 2) Olbe, L.; Carlsson, E.; Lindberg, P. A Proton-pump Inhibitor Expedition: the Case Histories of Omeprazole and S-omeprazole. Nature Reviews Drug Discovery 2003, 2, 132-139. 3) Jain, K. S.; Shah, A. K.; Bariwal, J.; Shelke, S. M.; Kale, A. P.; Jagtap, J. R.; Bhosale, A. V. Recent advances in proton pump inhibitors and management of acidpeptic disorders. Bioorg. Med. Chem. 2007, 15, 1181–1205. 4) Kahrilas, P. J. Gastroesophageal Reflux Disease. N Engl. J. Med. 2008, 359, 17001707 5) Najm, W. I. Peptic Ulcer Disease. Primary Care 2011, 38, 383-94 6) Milosavljevic, T.; Kostić-Milosavljević, M.; Jovanović, I.; Krstić M. Complications of Peptic Ulcer Disease. Dig. Dis. 2011, 29, 491–493. 7) Roberts, S.; McDonald, I. M. In Burger’s Medicinal Chemistry and Drug Discovery, 6th ed.; Abraham, D. J. Ed.; John Wiley: New Jersey, 2003; Vol. 42003, pp 86–121. 8) Wolff, M. M.; Soll, A. H. The Physiology of Gastric Acid Secretion. N. Engl. J. Med.

1988, 319, 1707-1715. 9) S. H.; Pang, Graham, D.Y. A clinical guide to using intravenous proton-pump inhibitors in reflux and peptic ulcers. Therap. Adv. Gastroenterol. 2010, 3, 11-22. 10) Marom, H.; Agranat, I. Racemization of the Gastrointestinal Antisecretory Chiral Drug S-omeprazole Magnesium via the Pyramidal Inversion Mechanism: A Theoretical Study. Chirality 2010, 22, 798–807. 11) Cancer, H.; Cheeseman, J. R.; Agranat, I. Conformational Spaces of the Gastrointestinal Antisecretory Chiral Drug Omeprazole: Stereochemistry and Tautomerism. Chirality 2006, 18, 10–16.

18 ACS Paragon Plus Environment

Page 18 of 24

Page 19 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

12) Andersson, T.; Hassan-Alin, M.; Hasselgren, G.; Röhss, K.; Weidolf, L. Pharmacokinetic Studies with S-omeprazole, the (S)-Isomer of Omeprazole. Clin. Pharmacokinet. 2001, 40, 411-426. 13) Anderson, T.; Roehss, K.; Bredberg, E.; Hassan-Alin, M. Pharmacokinetics and pharmacodynamics of S-omeprazole, the S-isomer of omeprazole. Aliment Pharmacol. Ther. 2001, 15, 1563-1569. 14) Andersson, T.; Weidolf, L. Stereoselective Disposition of Proton Pump Inhibitors. Clin. Drug Invest. 2008, 28, 263-279. 15) Shin, J. M.; Cho, Y. M.; Sachs, G. Chemistry of Covalent Inhibition of the Gastric (H+, K+)-ATPase by Proton Pump Inhibitors. J. Am. Chem. Soc. 2004, 126, 78007811. 16) Reyes‐Gonzáleza, J.; Gómezb, R. M.; Cortés‐Guzmán, F. Theoretical study of the smiles rearrangement in the activation mechanism of proton pump inhibitors. J. Phys. Org. Chem. 2012, 25 230–238. 17) Al-Matar, A. Kh.; El-Eswed, B.; Tutunji, M. F.; Kinetics of Acid Degradation of Proton Pump Inhibitors in the Presence of a Thiol. Int. J. Chem. Kinet. 2009, 41, 498506. 18) Brandstrom, A.; Bergman, N.-A.; Lindberg, P.; Grundevik, I.; Johansson, S.; Tekenbergs-Hjelte, L.; Ohlson, K. Chemical Reactions of Omeprazole and Omeprazole Analogues. II. Kinetics of the Reaction of Omeprazole in the Presence of 2-Mercaptoethanol. Acta Chem. Scand. 1989, 43, 549-568. 19) Brändström, A.; Lindberg, P.; Bergman, N.; Alminger, T.; Ankner, K.; Junggren, U.; Lamm, B.; Nordberg, P.; Erickson, M.; Grundevik, I. et al. Chemical Reactions of Omeprazole and Omeprazole Analogues. I. A Survey of the Chemical

19 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

Transformations of Omeprazole and its Analogues. Acta Chem. Scand. 1989, 43, 536548. 20) Brandstrom, A.; Bergman, N.-A.; Grundevik, I.; Johansson, S.; TekenbergsHjelte, L.; Ohlson, K. Chemical Reactions of Omeprazole and Omeprazole Analogues. III. Protolytic Behaviour of Compounds in the Omeprazole System. Acta Chem. Scand.

1989, 43, 569-576. 21) A. T. BRUNI, M. MIGUEL, C. FERREIRA, Theoretical Study of Omeprazole Behavior: Racemization Barrier and Decomposition Reaction. Int. J. Quantum Chem.

2008, 108, 1097–1106. 22) Ife, R. J.; Dyke, C. A.; Keeling, D. J.; Meenan, E.; Meeson, M. L.; Parsons, M. E.; Price,

C.

A.;

Theobald,

C.

J.;

Underwood,

A.

2-[[(4-Amino-2-

pyridyl)methyl]sulfinyl]benzimidazole H+/K+-ATPase inhibitors. The relationship between pyridine basicity, stability, and activity. J. Med. Chem. 1989, 32, 1970-1977. 23) Kromer, W.; Krüger, U.; Huber, R.; Hartmann, M.; Steinijans, V.W. Differences in pH-Dependent Activation Rates of Substituted Benzimidazoles and Biological in vitro Correlates. Pharmacology 1998, 56, 57–70. 24) Sachs, G.: Shin, J. M.; Vagin, O.; Lambrecht, N.; Yakubov, I.; Munson, K. The gastric H,K ATPase as a drug target: past, present, and future. J Clin Gastroenterol. 2007, 41, S226-S242. 25) Shin, J. M.; Munson, K.; Vagin, O.; Sachs, G. The gastric HK-ATPase: structure, function, and inhibition. Pflugers Arch. 2009, 457, 609-622. 26) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functional. Theor. Chem. Acc. 2008, 120, 215-241.

20 ACS Paragon Plus Environment

Page 21 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

27) Hariharan, P. C.; Pople, A. Accuracy of AH equilibrium geometries by single determinant molecular-orbital theory. Mol. Phys. 1974, 27, 209−214. 28) Hariharan, P. C.; Pople, J. A. Influence of polarization functions on molecular-orbital hydrogenation energies. Theor. Chem. Acc. 1973, 28, 213-222. 29) O’Reilly, R. J.; Karton, A.; Radom, L. Effect of substituents on the preferred modes of one-electron reductive cleavage of N−Cl and N−Br bonds. J. Phys. Chem. A 2013, 117, 460−472. 30) Yu, H.-Z.; Yang, Y.-M.; Zhang, L.; Dang, Z.-M.; Hu, G.-H. Quantum-Chemical predictions of pKa’s of thiols in DMSO. J. Phys. Chem. A 2014, 118, 606−622. 31) Fukui, K. The path of chemical-reactions - The IRC approach. Acc. Chem. Res. 1981, 14, 363-68. 32) Mineva, T.; Russo, N.; Toscano, M. Self Consistent Reaction Field Theory of Solvent Effects in the Framework of Gaussian Density Functional Method. Int. J. Quantum Chem. 1995, 56, 663-668. 33) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal solvation model based on solute electron density and a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 63786396. 34) Ghosh, D.; Sahu, D.; Saravanan, S.; Abdi, S. H. R.; Ganguly, B.; Khan, N. H.; Kureshy, R. I.; Bajaj, H. C. Synthetically amenable amide derivatives of tosylatedamino acids as organocatalysts for enantioselective allylation of aldehydes: computational rationale for enantioselectivity. Org. Biomol. Chem. 2013, 11, 3451– 3460.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

35) Saielli, G. Differential Solvation Free Energies of Oxonium and Ammonium Ions: Insights from Quantum Chemical Calculations. J. Phys. Chem. A 2010, 114, 7261– 7265. 36) Camaioni, D. M.; Schwerdtfeger, C. A. Comment on “Accurate Experimental Values for the Free Energies of Hydration of H+, OH-, and H3O+”. J. Phys. Chem. A 2005, 109, 10795–10797. 37) Thapa, B.; Schlegel, H. B. Density Functional Theory Calculation of pKa’s of Thiols in Aqueous Solution Using Explicit Water Molecules and the Polarizable Continuum Model. J. Phys. Chem. A 2016, 120, 5726−5735. 38) Magill, A. M.; Cavell, K. J.; Yates, B. F. Basicity of Nucleophilic Carbenes in Aqueous and Nonaqueous Solvents Theoretical Predictions. J. Am. Chem. Soc. 2004, 126, 8717-8724. 39) 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.; et al. Gaussian 09, Revision D.01; Gaussian, Inc:Wallingford, CT, 2010. 40) Schwamm, R. J.; Vianello, R.; Marsavelski, A.; García, M. Á.; Claramunt, R. M.; Alkorta, I.; Saame, J.; Leito, I.; Fitchett, C. M.; Edwards, A. J.; Coles, M. P. 15N NMR Spectroscopy, X-ray and Neutron Diffraction, Quantum Chemical Calculations, and UV/vis-Spectrophotometric Titrations as Complementary Techniques for the Analysis of Pyridine-Supported Bicyclic Guanidine Superbases. J. Org. Chem. 2016, 81, 7612-7625. 41) Xu, S.; Nilles, J. M.; Bowen, K. H. Zwitterion formation in hydrated amino acid, dipole bound anions: How many water molecules are required? J. Chem. Phys. 2003, 119, 10696-10701.

22 ACS Paragon Plus Environment

Page 22 of 24

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

42) Kendall, M. J. Review article: S-omeprazole - the first proton pump inhibitor to be developed as an isomer. Aliment Pharmacol. Ther. 2003, 17, 1–4. 43) Abelo, A.; Andersson, T.; Antonsson, M.; Naudot, A. K.; Skanberg, I.; Weidolf, L. Stereoselective metabolism of omeprazole by human cytochrome P450 enzymes. Drug Metab. Dispos. 2000, 28, 966–972. 44) Shirasaka, Y.; Sager, J. E.; Lutz, J. D.; Davis, C.; Isoherranen N. Inhibition of CYP2C19 and CYP3A4 by Omeprazole Metabolites and Their Contribution to DrugDrug Interactions. Drug Metab. Dispos. 2013, 41, 1414–1424.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Graphical Abstract

24 ACS Paragon Plus Environment

Page 24 of 24