1,255
Views
10
CrossRef citations to date
0
Altmetric
Research Article

Optimization of peptidyl allyl sulfones as clan CA cysteine protease inhibitors

, , , , , & show all
Pages 468-478 | Received 24 Nov 2011, Accepted 16 Dec 2011, Published online: 01 Mar 2012

Abstract

This research investigates the synthesis and inhibitory potency of a series of novel dipeptidyl allyl sulfones as clan CA cysteine protease inhibitors. The structure of the inhibitors consists of a R1-Phe-R2-AS-Ph scaffold (AS = allyl sulfone). R1 was varied with benzyloxycarbonyl, morpholinocarbonyl, or N-methylpiperazinocarbonyl substituents. R2 was varied with either Phe of Hfe residues. Synthesis involved preparation of vinyl sulfone analogues followed by isomerization to allyl sulfones using n-butyl lithium and t-butyl hydroperoxide. Sterics, temperature and base strength were all factors that affected the formation and stereochemistry of the allyl sulfone moiety. The inhibitors were assayed with three clan CA cysteine proteases (cruzain, cathepsin B and calpain I) as well as one serine protease (trypsin). The most potent inhibitor, (E)-Mu-Phe-Hfe-AS-Ph, displayed at least 10-fold selectivity for cruzain over clan CA cysteine proteases cathepsin B and calpain I with a kobs/[I] of 6080 ± 1390 M−1s−1.

Abbreviations
AMC,=

7-aminomethyl coumarin

AS,=

allyl sulfone

Cbz,=

benzyloxycarbonyl

DTT,=

dithiothreitol

EDCl,=

1-ethyl-3-(3-dimethylaminopropyl)carbodiimide

Elim,=

eliminated sulfone (-C(O)-N=C(R)-CH=CH2)

Hfe,=

homo-phenylalanine residue

HOBT,=

hydroxybenzotriazole

iBCF,=

isobutyl chloroformate

Mp,=

N-methylpiperazinocarbonyl

Mu,=

morpholinocarbonyl

NMM,=

N-methylmorpholine

VS,=

vinyl sulfone

Introduction

Proteases are a group of enzymes that are responsible for a variety of metabolic functions based on their ability to cleave peptide bondsCitation1–4. They are grouped by catalytic mechanism into five major classes: metallo, aspartate, threonine, serine and cysteine proteasesCitation5. Cysteine proteases catalyze the cleavage of peptide bonds using a catalytic dyad of Cys and His residuesCitation6. Each of the five major protease classes is further divided into clans and familiesCitation7. Clan CA (C indicates cysteine) is the largest of the cysteine protease clans and contains a variety of enzymes including calpains, cathepsins and cruzain.

Calpains are members of family C2, requiring Ca2+ for activationCitation8. They have been linked to a number of different cellular functions including signal transduction, cell fusion, mitosis and protein turnover. Mutations and other defects in calpains have been implicated in hypertension and muscular dystrophy as well as neurodegenerationCitation9. To date, no highly selective and potent inhibitors of calpains have been synthesized.

Cathepsins are members of family C1. Eleven cathepsins are currently known including cathepsin B, C, F, H, K (O2), L, O, S, V (L2), W and X (parentheses indicate secondary names). Cathepsin B is the most abundant of the cathepsins, serving a variety of cell degradation processesCitation10 in addition to potentially functioning as a defence mechanism against amyloid plaque aggregationCitation11. Cathepsin B has been associated with tumour progressionCitation6,Citation12,Citation13. More generally, cathepsins have been linked to rheumatoid arthritis, osteoarthritis, neurological disorders and lysosomal storage diseasesCitation14,Citation15. Because of their role in these diseases, cathepsins have been targets for drug developmentCitation16–22. Continued investigation of new cathepsin inhibitors provides a route for expanding treatment of these diseases.

Cruzain is another member of family C1. It is a key enzyme in the life cycle of Trypanosoma cruzi, the parasitic vector of Chagas diseaseCitation23–25. Chagas disease is an increasing problem in Latin America with 10 million people suffering from infection and millions more at riskCitation26. Current treatment utilizes benznidazole (a nitroimidazole derivative)Citation27 and nifurtimox (a 5-nitrofuran derivative)Citation28, which are effective in the acute phase of the disease; however, these treatments are plagued with severe side effects and increasing resistanceCitation29,Citation30. As a result, demand for developing new treatments for this disease remains prevalent.

To date, a number of cysteine protease inhibitors have been developed using a variety of reactive functional groupsCitation31, also referred to as “warheads,” including aldehydesCitation20, fluoromethyl ketonesCitation21, nitrilesCitation32, α-ketoesters, amides, and acidsCitation33, tetrafluorophenoxymethyl ketonesCitation34, epoxysuccinatesCitation35, thiosemicarbazonesCitation36, hydrazone derivativesCitation37,Citation38, vinyl sulfonesCitation39 and allyl sulfonesCitation40. These warheads are attached to either a peptidyl or peptidomimetic scaffoldCitation16,Citation41.

We selected cruzain as our primary target due to the high demand for the development of new treatments for Chagas disease. Currently, the most effective known inhibitor of cruzain, referred to as K11777, is a dipeptidyl vinyl sulfone (Mp-Phe-Hfe-VS-Ph)Citation42. Both metabolicCitation43 and animal studiesCitation44,Citation45 have confirmed that K11777 is a highly selective and potent inhibitor of cruzain in vivo and is currently entering clinical trialsCitation46. We have demonstrated previously that use of an allyl sulfone in place of the vinyl sulfone moiety showed enhancements in inhibitory potency for clan CA cysteine proteases ()Citation40.

Figure 1.  Basic scaffold of dipeptidyl allyl sulfone inhibitors with enzyme subsite designations P3, P2, P1 and P1′.

Figure 1.  Basic scaffold of dipeptidyl allyl sulfone inhibitors with enzyme subsite designations P3, P2, P1 and P1′.

In order to optimize the lead allyl sulfones for inhibition of cruzain, we chose to modify the P3 and P1 positions of the scaffold. The peptidyl scaffold was maintained because no peptidomimetic scaffolds have been able to achieve in vivo potency equivalent to K11777. The goal of the P3 modification was to vary solubility and polarity inside the active site using Cbz, Mp and Mu groups. Based on previous kinetic data, the P2 and P1′ positions were kept constant, as Phe and SO2Ph, respectively, in all analoguesCitation39. The P1 residue was varied, using either Phe or Hfe residues, in order to examine the effect of flexibility of the phenyl group after the introduction of the planar allyl sulfone moiety.

We have previously synthesized allyl sulfones from vinyl sulfone precursors and determined that stereochemistry of the P1 residue of the vinyl sulfone influenced the stereochemistry of the allyl sulfone product, although the contributing effect was not determined. To further investigate this, both enantiomers of the P1 amino acid residue were used during the synthesis of the dipeptidyl allyl sulfones.

Materials and methods

Materials

Starting materials and reagents were purchased from VWR International (Radnor, Pennsylvania) and used without further purification. Amino acids were purchased from Advanced ChemTech (Louisville, Kentucky) and enzymes and substrates from Calbiochem (San Diego, California) and Enzo Life Sciences (Ann Arbor, Michigan). Compound purity was confirmed by MS electron spray ionization (ESI) and NMR (13C and 1H with chemical shifts reported in ppm relative tetramethyl silane). MS analysis was performed at the University of Arkansas High Performance Mass Spectrometry Laboratory using a Bruker Esquire-LC Ion Trap LC/MS. NMR experiments were conducted using a BrukerAvance III 400 MHz UltrashieldPlus Spectrometer. Compounds were purified via gravity column chromatography using Sigma-Aldrich Fluka analytical silica (60 Å, 70–230 mesh). Reaction progress was monitored using EMD 60 F254 TLC plates. Fluorogenic enzyme assays were performed on a BioTek Flx800 Plate Reader. Molecular modelling data was acquired using chemical computing group’s Molecular Operating Environment (MOE) software package.

General synthesis of inhibitors

Trans-vinyl sulfones 1–4 were prepared as previously described via Wadsworth-Emmons chemistry using chiral amino acid aldehydes and a sulfonylphosphonateCitation40,Citation47. Phe-OMe was reacted with morpholinocarbonyl chloride and saponified to carboxylic acid 5 (Mu-Phe-OH) in 71% overall yield (). Phe-OBz was reacted with N-methylpiperazinocarbonyl chloride and hydrogenated to give carboxylic acid 6 (Mp-Phe-OH) in 89% yield. trifluoroacetic acid (TFA) salts 1 and 2 (TFA·L/D-Hfe-VS-Ph) were coupled to 5 using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDCl) and hydroxybenzotriazole (HOBT) to produce dipeptidyl vinyl sulfones 7 and 8 (Mu-Phe-L/D-Hfe-VS-Ph). TFA salts 3 and 4 (TFA·L/D-Phe-VS-Ph) were coupled to Cbz-Phe-OH using EDCl and HOBT to yield dipeptidyl vinyl sulfones 9 and 10 (Cbz-Phe-L/D-Phe-VS-Ph). TFA salts 1–3 were coupled to 6 using iBCF and N-methylmorpholine (NMM) to give dipeptidyl vinyl sulfones 11–13 (Mp-Phe-L/D-Hfe-VS-Ph, Mp-Phe-L-Phe-VS-Ph). Dipeptidyl vinyl sulfones 7–13 were then isomerized to allyl sulfone analogues 14–20 using t-BuOOH and n-BuLi.

Scheme 1.  (a) EDCl, HOBT, Et3N, DCM, overnight; (b) iBCF, NMM, THF/DMF, overnight; (c) 1 eq.n-BuLi, 4.5 eq. t-BuOOH, −20°C for 70 min, 9 h or 18 h.

Scheme 1.  (a) EDCl, HOBT, Et3N, DCM, overnight; (b) iBCF, NMM, THF/DMF, overnight; (c) 1 eq.n-BuLi, 4.5 eq. t-BuOOH, −20°C for 70 min, 9 h or 18 h.

Detailed synthesis

Preparation of Amino Acid Vinyl Sulfone TFA Salts

TFA salts 1–4 were synthesized using previously reported methodsCitation40.

Synthesis of (2S)-2-(morpholine-4-carboxamido)-3-phenylpropanoic acid (5, Mu-Phe-OH) 

The hydrochloride salt of Phe-OMe (7.00 g, 32.5 mmol) was dissolved in THF (150 mL) under argon at −15°C. NMM (7.25 g, 71.4 mmol) was added followed by the dropwise addition of morpholinecarbonylchloride (5.34 g, 35.7 mmol). The reaction was allowed to warm to room temperature and stirred overnight. The reaction was poured into aq. HCl (1 N, 100 mL) and extracted with EtOAc (3 × 60 mL). The organic extract was then washed with aq. HCl (1 N, 3 × 60 mL) and brine (3 × 60 mL) and dried (MgSO4). The solvent was removed under reduced pressure. The product was initially isolated as a syrup that spontaneously crystallized into white crystals with a 92% yield. 1H NMR (400 MHz, CDCl3) δ 3.14 (t, 2H, CHCH2Ph), 3.33 (m, 4H, CH2NCH2), 3.75 (t, 4H, CH2OCH2), 4.81 (m, 1H, α-H), 4.87 (d, 1H, NH), 7.11–7.12 (m, 2H, Ph), 7.28–7.32 (m, 3H, Ph). Mu-Phe-OMe (9.47 g, 32.5 mmol) was dissolved in methanol (150 mL). Sodium hydroxide (2 N, 2.62 g, 48.5 mmol) was added. After 4 h, the solvent was removed under reduced pressure and the mixture was suspended in water (100 mL). The solution was acidified to pH 2 by addition of aq. HCl (3 N, approx. 20 mL). The product was extracted with EtOAc (3 × 30 mL), washed with brine (2 × 25 mL) and dried (MgSO4). The solvent was removed under reduced pressure to give 5 as a white powder in a 76.5% yield. 1H NMR (400 MHz, DMSO-d6) δ 2.90 (dd, 1H, CH2Ph J = 13 Hz, 10 Hz), 3.03 (dd, 1H, CH2Ph, J = 15 Hz, 5 Hz), 3.23 (m, 4H, CH2NCH2), 3.49 (m, 4H, CH2OCH2), 4.23 (m, 1H, α-H), 6.73 (d, 1H, NH), 7.17–7.30 (m, 5H, Ph), 12.53 (s, broad, 1H, COOH).

Synthesis of (2S)-2-(N-methylpiperazine-4-carboxamido)-3-phenylpropanoic acid (6, Mp-Phe-OH) 

The hydrochloride salt of N-methylpiperazinecarbonyl chloride (1.50 g, 7.54 mmol) was dissolved in THF (50 mL) under argon at −15°C. Et3N (0.839 g, 8.29 mmol) was added and the reaction was stirred for 15 min. The hydrochloride salt of Phe-OBz (2.00 g, 6.84 mmol) was dissolved in THF (50 mL) under argon at −15°C. Et3N (0.839 g, 8.29 mmol) was added and stirred for 15 min. The two reactions were combined and kept at −15°C for 15 min, then allowed to warm to room temperature and stirred overnight. The reaction was quenched with water (80 mL) and the product was extracted with EtOAc (3 × 50 mL). The organic layer was washed with water (3 × 50 mL) and brine (2 × 50 mL) and dried (MgSO4). The solvent was removed under reduced pressure. The crude product was purified via column chromatography (silica, 5% MeOH in DCM, 0.5% Et3N). The pure product (Mp-Phe-OBz) was isolated as a clear colorless syrup with a yield of 88.5%. 1H NMR (400 MHz, CDCl3) δ 2.31 (s, 3H, CH3N), 2.37 (t, 4H, CH2N(CH3)CH2), 3.14 (m, 2H, CH2Ph), 3.36 (m, 4H, CH2NCH2), 3.88 (broad, 2H, α-H and NH), 7.00–7.03 (m, 2H, Ph), 7.21–7.41 (m, 8H, Ph). Mp-Phe-OBz was deprotected using a procedure adapted from Maring et al.Citation48. Mp-Phe-OBz (0.600 g, 1.52 mmol) was dissolved in ethanol (100 mL) and degassed. Ammonium formate (0.957 g, 15.171 mmol) and Pd/C (120 mg) were added. The reaction was refluxed for 40 min. The resulting solution was filtered over Celite and the solvent was removed under reduced pressure. The product was isolated as a white powder with a yield of 85.3%. 1H NMR (400 MHz, CDCl3) δ 2.37 (s, 3H, CH3N), 2.54 (broad, 4H, CH2N(CH3)CH2), 3.01 (m, 1H, CH2Ph), 3.18 (m, 1H, CH2Ph), 3.34–3.45 (m, 4H, CH2NCH2), 4.44 (m, 1H, α-H), 5.73 (d, 1H, NH), 7.12–7.26 (m, 5H, Ph). Acid peak not observed.

General procedure for peptide coupling reaction using Mu-Phe-OH. Phenyl (3S)-3-(N-morpholine-4-carboxamidophenylalanyl)amino-5-phenylpent-1-enyl sulfone (7, Mu-Phe-L-Hfe-VS-Ph) 

TFA salt 1 (TFA·D-Hfe-VS-Ph) (2.89 g, 6.95 mmol) was dissolved in DCM (120 mL) and Et3N (1.06 g, 10.4 mmol) was added. Compound 5 (Mu-Phe-OH) (1.99 g, 6.95 mmol) was dissolved in DCM (80 mL) and added to the reaction followed by HOBT (1.03 g, 7.64 mmol) and EDCl (1.19 g, 7.64 mmol). Additional Et3N was added to adjust the pH to between 9 and 10. The reaction was stirred overnight. The mixture was washed with aq. HCl (3N, 3 × 50 mL), aq. NaHCO3 (10%, 3 × 50 mL), brine (3 × 50 mL) and dried (MgSO4). The solvent was removed under reduced pressure. The product was purified via column chromatography (silica, 1% MeOH in DCM) and isolated as a white powder with an 86.1% yield. 1H NMR (400 MHz, CDCl3) δ 1.79–1.89 (m, 2H, CH2CH2Ph), 2.57 (m, 2H, CH2CH2Ph), 3.05 (m, 2H, CH2Ph), 3.28 (m, 4H, CH2NCH2), 3.62 (m, 4H, CH2OCH2), 4.47 (q, 1H, α-H), 4.62 (s, broad, 1H, α-H), 4.96 (m, 1H, NH), 6.10 (dd, 1H, CH=CH-SO2, J = 15 Hz, 2 Hz), 6.39 (d, 1H, NH), 6.80 (dd, 1H, CH=CH-SO2, J = 15 Hz, 8 Hz), 7.01–7.33 (m, 10H, 2 × Ph), 7.50–7.91 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 31.71, 35.68, 38.14, 43.96, 49.13, 56.19, 66.31, 126.35, 127.31, 127.68, 128.33, 128.58, 128.62, 128.92, 129.15, 129.21, 129.30, 130.65, 133.50, 136.62, 145.26, 157.13, 171.51.

Synthesis of phenyl (3R)-3-(N-morpholine-4-carboxamidophenylalanyl)amino-5-phenylpent-1-enyl sulfone. (8, Mu-Phe-D-Hfe-VS-Ph) 

TFA salt 2 (TFA·L-Hfe-VS-Ph) was used as the starting material and the general peptide coupling procedure was followed. The product was purified via recrystallization (1:3 Hex: EtOAc) and isolated as white crystals with an 84.6% yield. 1H NMR (400 MHz, CDCl3) δ 1.68–1.82 (m, 2H, CH2CH2Ph), 2.41 (t, 2H, CH2CH2Ph), 3.07 (m, 2H, CH2Ph), 3.25 (m, 4H, CH2NCH2), 3.63 (m, 4H, CH2OCH2), 4.42 (q, 1H, α-H), 4.62 (s, broad, 1H, α-H), 4.88 (d, 1H, NH), 6.18 (d, 1H, NH), 6.52 (d, 1H, CH=CH-SO2), 6.79 (dd, 1H, CH=CH-SO2, J = 15 Hz, 5 Hz) 7.05–7.37 (m, 10H, 2 × Ph), 7.50–7.93 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 31.69, 35.49, 37.86, 43.89, 49.20, 56.25, 66.29, 126.31, 127.21, 127.64, 128.34, 128.57, 128.84, 129.21, 129.29, 130.73, 133.44, 136.67, 140.26, 140.32, 145.42, 157.21, 171.54.

General procedure for peptide coupling reaction using Cbz-Phe-OH. Phenyl (3S)-3-(N-carbobenzyloxyphenylalanyl)amino-4-phenylbut-1-enyl sulfone (9, Cbz-Phe-L-Phe-VS-Ph) 

TFA salt 3 (TFA·L-Phe-VS-Ph) (0.946 g, 2.36 mmol) was dissolved in DCM (75 mL). Et3N (0.572 g, 2.76 mmol) was added, followed by Cbz-Phe-OH (0.776 g, 2.59 mmol), EDCl (0.497 g, 2.59 mmol) and HOBT (0.350 g, 2.59 mmol). Additional Et3N was added to adjust the pH to between 9 and 10. The reaction was stirred overnight. The reaction was washed with citric acid (10%, 3 × 25 mL), aq. NaHCO3 (10%, 3 × 25 mL), brine (2 × 25 mL) and dried (MgSO4). The solvent was removed under reduced pressure. The product was purified via column chromatography (silica, 1:1 Hex: EtOAc) as a white powder with a 36.4% yield. 1H NMR (400 MHz, CDCl3) δ 2.81 (d, 2H, CH2Ph), 2.95–3.05 (m, 2H, CH2Ph), 4.29 (q, 1H, α-H), 4.93 (q, 1H, α-H), 5.07 (s, 2H, OCH2Ph), 5.14 (m, 1H, NH), 5.73 (d, 1H, NH), 5.95 (d, 1H, CH=CH-SO2), 6.80 (dd, 1H, CH=CH-SO2, J = 15 Hz, 5 Hz), 6.99–7.46 (m, 15H, 3 × Ph), 7.52–7.83 (m, 5H, Ph-SO2).

Synthesis of phenyl (3R)-3-(N-carbobenzyloxyphenylalanyl)amino-4-phenylbut-1-enyl sulfone (10, Cbz-Phe-D-Phe-VS-Ph) 

See the above procedure for peptide coupling reaction using Cbz-Phe-OH. TFA salt 4 (TFA·D-Phe-VS-Ph) was used as starting material. The product was purified via column chromatography (silica, 1:1 Hex: EtOAc) as a white powder with a 64.1% yield. 1H NMR (400 MHz, CDCl3) δ 2.61–3.14 (m, 4H, 2 × CH2Ph), 4.33 (q, 1H, α-H), 4.89–5.15 (m, 3H, α-H and CH2OPh), 5.34 (m, 1H, NH), 6.19 (d, 1H, NH), 6.32 (d, 1H, CH=CH-SO2), 6.80–7.44 (m, 16 H, 2 × CH2Ph, OPh, CH=CH-SO2), 7.47–7.87 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 38.16, 39.90, 40.16, 50.27, 56.54, 67.22, 127.21, 127.26, 127.67, 128.07, 128.11, 128.32, 128.60, 128.71, 128.81, 129.16, 129.27, 129.31, 131.15, 133.48, 135.28, 136.06, 140.03, 144.75, 170.48.

General procedure for peptide coupling using Mp-Phe-OH. Phenyl (3S)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylpent-1-enyl sulfone (11, Mp-Phe-L-Hfe-VS-Ph) 

Compound 6 (Mp-Phe-OH) (0.874 g, 1.23 mmol) was dissolved in 3:1 THF/DMF (40 mL) at −10°C. NMM (0.137 g, 1.35 mmol) was added and the reaction was stirred for 20 min with the gradual addition of iBCF (0.185 g, 1.35 mmol). The reaction was stirred for 1 h maintaining a temp below −5°C. TFA salt 1 (TFA·L-Hfe-VS-Ph) was added at −10°C. Additional NMM (0.137g, 1.35 mmol) was added and the reaction was stirred for 1 h at −10°C. The reaction was raised to room temperature and stirred overnight. Aq. NaHCO3 (10%, 200 mL) was added and the product was extracted using EtOAc (3 × 40 mL). The organic layers were combined and washed with brine (2 × 50 mL) and dried (MgSO4). The solvent was removed under reduced pressure and the crude product was purified by column chromatography (silica, 5% MeOH in DCM, 0.5% Et3N). The product was isolated as a clear colorless syrup with a 32.2% yield. 1H NMR (400 MHz, CDCl3) δ 1.74 (m, 1H, CH2CH2Ph), 1.85 (m, 1H, CH2CH2Ph), 2.25 (s, 3H, CH3N), 2.31 (m, 4H, CH2N(CH3)CH2), 2.54 (m, 2H, CH2CH2Ph), 3.00 (m, 2H, CH2Ph), 3.28 (m, 4H, CH2NCH2), 4.58 (broad, 2H, α-H, α-H), 5.23 (d, 1H, NH), 6.10 (dd, 1H, CH=CH-SO2, J = 15 Hz, 2 Hz), 6.77 (dd, 1H, CH=CH-SO2, J = 15 Hz, 8 Hz), 7.01–7.33 (m, 11H, 2 × Ph, NH), 7.50–7.91 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 31.77, 35.67, 38.57, 43.73, 46.02, 49.07, 54.45, 55.96, 126.22, 127.08, 127.63, 128.38, 128.52, 128.64, 129.28, 129.30, 130.32, 133.48, 136.67, 140.18, 140.47, 145.83, 156.95, 172.01. MS (ESI) m/z 575.2 [M+H]+.

Synthesis of phenyl (3R)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylpent-1-enyl sulfone (12, Mp-Phe-D-Hfe-VS-Ph) 

See general procedure for peptide coupling using Mp-Phe-OH. TFA salt 2 (TFA·D-Hfe-VS-Ph) was used as the starting material. The crude product was purified using column chromatography (silica, 5% MeOH in DCM, 0.5% Et3N). The product was isolated as a clear colorless syrup with a 22% yield. 1H NMR (400 MHz, CDCl3) δ 1.68 (m, 1H, CH2CH2Ph), 1.80 (m, 1H, CH2CH2Ph), 2.32 (s, 3H, CH3N), 2.41 (m, 6H, CH2N(CH3)CH2 and CH2CH2Ph), 3.07 (m, 2H, CH2Ph), 3.32 (m, 4H, CH2NCH2), 4.41 (m, 1H, α-H), 4.62 (m, 1H, α-H), 4.91 (d, 1H, NH), 6.33 (d, 1H, NH) 6.52 (dd, 1H, CH=CH-SO2, J = 15 Hz, 2 Hz), 6.85 (dd, 1H, CH=CH-SO2,J = 15 Hz, 8 Hz), 7.01–7.33 (m, 10H, 2 × Ph), 7.50–7.91 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 31.7, 35.52, 37.76, 43.65, 46.00, 49.17, 54.41, 56.28, 60.41, 126.27, 127.16, 127.64, 128.36, 128.56, 128.82, 129.23, 129.28, 130.68, 133.43, 136.76, 140.37, 145.5, 157.04, 171.68.

Synthesis of phenyl (3S)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylbut-1-enyl sulfone (13, Mp-Phe-L-Phe-VS-Ph) 

See general procedure for peptide coupling using Mp-Phe-OH. TFA salt 3 (TFA·L-Phe-VS-Ph) was used as the starting material. The crude product was purified using column chromatography (silica, 5% MeOH in DCM, 0.5% Et3N). The product was isolated as a clear colorless syrup with a 45.1% yield. 1H NMR (400 MHz, CDCl3) δ 2.33 (m, 7H, CH2N(CH3)CH2 and CH2N(CH3)CH2), 2.85 (dd, 2H, CH2Ph, J =  6 Hz and 3 Hz), 3.01 (m, 2H, CH2Ph), 3.28 (m, 4H, CH2NCH2), 4.41 (q, 1H, α-H), 4.76 (d, 1H, NH), 4.93 (m, 1H, α-H), 6.03 (dd, 1H, CH=CH-SO2, J = 15 Hz, 2Hz), 6.22 (d, 1H, NH), 6.84 (dd, 1H, CH=CH-SO2), J = 15 Hz, 5 Hz), 7.01–7.30 (m, 10H, 2 × Ph), 7.50–7.83 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 37.97, 40.17, 43.67, 54.48, 55.89, 127.07, 127.22, 127.68, 128.62, 128.68, 128.87, 129.16, 129.24, 129.29, 129.35, 130.86, 133.44, 135.51, 136.73, 140.11, 144.76, 156.78, 171.36.

Preparation of allyl sulfones: General procedure for vinyl sulfone isomerization. Phenyl (E)-3-(N-morpholine-4-carboxamidophenylalanyl)amino-5-phenylpent-2-enyl sulfone (14, E-Mu-Phe-Hfe-AS-Ph) 

To dry THF (40 mL), t-BuOOH (4.01 mmol, 1.2 mL, 3.3 M in toluene) was added dropwise under argon at −78°C, followed by n-BuLi (1.02 mmol, 0.65 mL, 1.6 M in hexane). Compound 7 (Mu-Phe-L-Hfe-VS-Ph) (0.500 g, 0.890 mmol) was dissolved in dry THF (5 mL) and added dropwise. The reaction was allowed to warm to −20°C and stirred for 9 h. The reaction was quenched with aq. NH4Cl (10%, 50 mL) and the aqueous layer was extracted with EtOAc (3 × 50 mL). The organic layers were combined and washed with water (3 × 50 mL), brine (2 × 50 mL) and dried (MgSO4). The solvent was removed under reduced pressure and the crude product was purified via column chromatography (silica, 1% MeOH in DCM). The product was isolated as a white powder with a 6.7% yield. 1H NMR (400 MHz, CDCl3) δ 2.71 (t, 2H, CH2CH2Ph), 2.87 (t, 2H, CH2CH2Ph), 3.21 (m, 2H, CH2Ph), 3.37 (m, 4H, CH2NCH2), 3.60 (dd, 2H, CH2SO2, J = 2 Hz, 8 Hz), 3.67 (m, 4H, CH2OCH2), 4.64 (m, 2H, α-H and C=CH), 4.84 (d, 1H, NH), 7.13–7.40 (m, 10H, 2 × Ph), 7.45–7.66 (m, 5H, Ph-SO2), 8.67 (s, broad, 1H, NH). 13C NMR (100 MHz, CDCl3) δ 33.71, 35.77, 37.72, 43.79, 54.47, 56.29, 66.30, 103.83, 126.00, 127.41, 128.27, 128.43, 128.63, 129.07, 129.13, 129.28, 133.83, 136.59, 138.33, 141.15, 145.45, 156.89, 171.05. MS (ESI) m/z 562.2 [M+H]+.

Synthesis of phenyl (Z)-3-(N-morpholine-4-carboxamidophenylalanyl)amino-5-phenylpent-2-enyl sulfone (15, Z-Mu-Phe-Hfe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 8 (Mu-Phe-D-Hfe-VS-Ph) was used as the starting material. The crude product was purified via column chromatography (silica, 1% MeOH in DCM) and isolated as a white powder with an 8.6% yield. 1H NMR (400 MHz, CDCl3) δ 2.71 (t, 2H, CH2CH2Ph), 2.87 (t, 2H, CH2CH2Ph), 3.21 (m, 2H, CH2Ph), 3.37 (m, 4H, CH2NCH2), 3.60 (dd, 2H, CH2SO2, J = 1 Hz, 8 Hz), 3.67 (m, 4H, CH2OCH2), 4.64 (m, 2H, α-H and C=CH), 4.84 (d, 1H, NH), 7.13–7.40 (m, 10H, 2 × Ph), 7.45–7.66 (m, 5H, Ph-SO2), 8.67 (s, broad, 1H, NH). 13C NMR (100 MHz, CDCl3) δ 33.71, 35.77, 37.72, 43.79, 54.47, 56.29, 66.30, 103.83, 126.00, 127.41, 128.27, 128.43, 128.63, 129.07, 129.13, 129.28, 133.83, 136.59, 138.33, 141.15, 145.45, 156.90, 171.06. MS (ESI) m/z 562.2 [M+H]+.

Synthesis of phenyl (E)-3-(N-carbobenzyloxyphenylalanyl)amino-4-phenylbut-2-enyl sulfone (16, E-Cbz-Phe-Phe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 9 (Cbz-Phe-L-Phe-VS-Ph) was used as the starting material. The reaction was stirred for 70 min at −20°C. The crude product was purified via column chromatography (silica, 2:1 Hex: EtOAc). The final compound was isolated as a white powder with a 15.8% yield. 1H NMR (400 MHz, CDCl3) δ 3.03 (m, 2H, CH2Ph), 3.67–3.89 (m, 4H, PhCH2C=CHCH2SO2), 4.35 (d, 1H, α-H), 4.83 (t, 1H, C=CH), 5.12 (s, broad, 3H, NH, OCH2Ph), 7.09–7.39 (m, 15H, 3 × Ph), 7.45–7.75 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 37.75, 40.14, 54.47, 56.77, 67.40, 106.61, 126.76, 127.29, 128.24, 128.26, 128.29, 128.48, 128.53, 128.98, 129.16, 129.17, 129.24, 133.93, 135.95, 135.97, 137.36, 138.31, 144.45, 156.02, 170.11.

Synthesis of phenyl (Z)-3-(N-carbobenzyloxyphenylalanyl)amino-4-phenylbut-2-enyl sulfone (17, Z-Cbz-Phe-Phe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 10 (Cbz-Phe-D-Phe-VS-Ph) was used as the starting material. The reaction was stirred for 70 min at −20°C. The product was purified by column chromatography (silica, 2:1 Hex: EtOAc) and isolated as a white powder with a 16.7% yield. 1H NMR (400 MHz, CDCl3) δ 3.03 (m, 2H, CH2Ph), 3.67–3.89 (m, 4H, PhCH2C=CHCH2SO2), 4.35 (d, 1H, α-H), 4.83 (t, 1H, C=CH), 5.12 (s, broad, 3H, NH, OCH2Ph), 7.09–7.39 (m, 15H, 3 × Ph), 7.45–7.75 (m, 5H, Ph-SO2). 13C NMR (100 MHz, CDCl3) δ 37.76, 40.16, 54.49, 56.80, 67.38, 106.67, 126.77, 127.28, 128.24, 128.29, 128.48, 128.54, 128.60, 128.62, 128.97, 129.16, 129.18, 129.25, 133.94, 136.00, 137.37, 138.31, 144.38, 156.07, 170.1.

Synthesis of phenyl (E)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylpent-2-enyl sulfone (18, E-Mp-Phe-Hfe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 11 (Mp-Phe-L-Hfe-VS-Ph) was used as the starting material. The reaction was stirred for 8 h at −20°C and stored at −20°C for an additional 12 h. The crude product was purified via column chromatography (silica, 3% MeOH in DCM, 0.5% Et3N) and isolated as a white powder with a yield of 8.2%. 1H NMR (400 MHz, CDCl3) δ 2.29 (s, 3H, CH2N(CH3)CH2), 2.37 (m, 4H, CH2N(CH3)CH2), 2.69 (t, 3H, CH2CH2Ph), 2.83 (t, 3H, CH2CH2Ph), 3.18 (m, 2H, CH2Ph), 3.41 (m. 4H, CH2NCH2), 3.59 (m, 2H, CH2-SO2), 4.61 (q, 1H, α-H), 4.67 (t, 1H, C=CH), 4.86 (d, 1H, NH), 7.13–7.40 (m, 10H, 2 × Ph), 7.44–7.68 (m, 5H, Ph-SO2), 8.61 (s, 1H, NH).Citation13C NMR (100 MHz, CDCl3) δ 33.73, 35.88, 37.72, 43.62, 46.06, 54.44, 54.49, 56.37, 103.91, 125.97, 127.32, 128.30, 128.42, 128.62, 129.01, 129.10, 129.32, 133.78, 136.68, 138.44, 141.16, 145.20, 156.84, 171.13. MS (ESI) m/z 575.2 [M+H]+.

Synthesis of phenyl (Z)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylpent-2-enyl sulfone (19, Z-Mp-Phe-Hfe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 12 (Mp-Phe-D-Hfe-VS-Ph) was used as the starting material. The reaction was stirred for 8 h at −20°C and stored at −20°C for an additional 12 h. The crude product was purified via column chromatography (silica, 5% MeOH in DCM, 0.5% Et3N) and isolated as a white powder with a 6.7% yield. 1H NMR (400 MHz, CDCl3) δ 2.28 (s, 3H, CH2N(CH3)CH2), 2.37 (m, 4H, CH2N(CH3)CH2), 2.68 (t, 3H, CH2CH2Ph), 2.83 (t, 3H, CH2CH2Ph), 3.18 (m, 2H, CH2Ph), 3.39 (m, 4H, CH2NCH2), 3.59 (m, 2H, CH2-SO2), 4.61 (q, 1H, α-H), 4.67 (t, 1H, C=CH), 4.86 (d, 1H, NH), 7.13–7.40 (m, 10H, 2 × Ph), 7.44–7.68 (m, 5H, Ph-SO2), 8.61 (s, 1H, NH). 13C NMR (100 MHz, CDCl3) δ 33.74, 35.88, 37.71, 43.62, 45.92, 53.44, 54.44, 56.37, 103.89, 125.97, 127.32, 128.30, 128.42, 128.62, 129.01, 129.10, 129.32, 133.78, 136.68, 138.45, 141.16, 145.20, 156.84, 171.12. MS (ESI) m/z 575.2 [M+H]+.

Synthesis of phenyl (E)-3-(N-(N-methylpiperazine)-4-carboxamidophenylalanyl)amino-5-phenylbut-2-enyl sulfone (20, E-Mp-Phe-L-Phe-AS-Ph) 

See above procedure for vinyl sulfone isomerization. Compound 13 (Mp-Phe-L-Phe-VS-Ph) was used as the starting material. The reaction was stirred for 3 h at −20°C. The product was purified via column chromatography (silica, 3% MeOH in DCM, 0.5% Et3N) and isolated as a white powder with a yield of 7.2%. 1H NMR (400 MHz, CDCl3) δ 2.27 (s, 3H, CH2N(CH3)CH2), 2.34 (m, 4H, CH2N(CH3)CH2), 2.93–3.18 (m, 2H, CH2Ph), 3.32 (m, 4H, CH2NCH2), 3.61–3.91 (m, 4H, CH2-SO2 and CH2Ph), 4.46 (q, 1H, α-H), 4.79 (m, 2H, α-H and NH), 7.06–7.41 (m, 10H, 2 × Ph), 7.46–7.85 (m, 5H, Ph-SO2), 8.45 (s, 1H, NH). 13C NMR (100 MHz, CDCl3) δ 37.58, 40.20, 43.60, 52.93, 54.40, 54.63, 56.25, 105.78, 126.65, 127.16, 128.29, 128.42, 128.89, 129.12, 129.21, 129.24, 133.84, 136.64, 137.57, 138.45, 144.39, 156.83, 171.15. MS (ESI) m/z 561.2 [M+H]+.

Procedure for elimination of Mu-Phe-Hfe-VS-Ph. (2S)-N-(5-phenylpent-1-en-3-ylidene)-2-(N-morpholine-4-carboxamidophenylalanyl)amino-3-phenylpropanamide. (21, Mu-Phe-Hfe-Elim) 

See general procedure for isomerization of the vinyl sulfone. Compound 7 was used as the starting material. The reaction was allowed to warm to room temperature and instead of adding both n-BuLi and t-BuOOH at −78°C, only n-BuLi was added. The product was purified by column chromatography (silica, 3% MeOH in DCM) and isolated as a white powder with a 6.4% yield. 1H NMR (400 MHz, CDCl3) δ 2.13 (m, 1H, CH2CH2Ph), 2.44 (m, 2H, 1H of CH2CH2Ph, 1H of CH2CH2Ph), 2.63 (m, 1H, CH2CH2Ph), 2.92 (m, 2H, CH2NCH2), 3.11 (m, 1H, CH2Ph), 3.26 (m, 3H, 1H of CH2Ph, 2H, of CH2NCH2), 3.52 (m, 2H, CH2OCH2), 3.63 (m, 2H, CH2OCH2), 4.50 (t, 1H, α-H), 5.02 (m, 2H, CH=CH2), 6.11 (dd, 1H, CH=CH2, J = 19 Hz, 11 Hz), 7.10–7.35 (m, 10H, 2 × Ph), 7.60 (s, broad, 1H, NH). 13C NMR (100 MHz, CDCl3) δ 29.60, 37.41, 37.71, 46.80, 61.97, 66.38, 78.62, 114.24, 126.25, 127.02, 128.32, 128.60, 129.77, 129.82, 136.55, 140.60, 140.97, 159.09, 172.24. MS (ESI) m/z 442.2 [M+Na]+.

Enzyme assays

General assay parameters

All assays used the incubation method to measure the irreversible inhibition of the specified enzymeCitation49. Enzyme activity was observed via monitoring hydrolysis of various fluorogenic 7-amino-4-methylcoumarin (AMC) substrates using a BioTek Flx800 Plate Reader (λex = 360 nm, λem = 460 nm). Enzyme activity was measured at the specified incubation temperature. All assay conditions contained less than 8.3% dimethylsulfoxide (DMSO). Pseudo first-order inactivation rate constants were obtained from ln vt/v0 versus time plots.

Cathepsin B AssayCitation40

Human liver cathepsin B (0.259 mg/mL) was purchased from Calbiochem. A D-10 dilution of cathepsin B was prepared with a 5 µL aliquot of the stock enzyme and 45 µL of deionized water. A D-100 activation dilution was prepared by combining 267 µL kinetic buffer (0.1 M potassium phosphate, 1.25 mM EDTA, 0.01% Brij 35 at pH 6.0), 30 µL of the D-10 solution and 3 µL of 0.1 M dithiothreitol (DTT) (freshly prepared). The D-100 activation dilution was allowed to activate for 1 h at 0°C. The incubation solution contained 30 µL of inhibitor stock (10 µM-5 mM in DMSO), 300 µL kinetic buffer and 30 µL of D-100 enzyme prep at 25°C. Aliquots (50 µL) of the incubation solution were withdrawn at varying intervals and combined with 200 µL of substrate buffer (500 µM Cbz-Arg-Arg-AMC [Calbiochem Cathepsin B Substrate III], 0.25% DMSO v/v, 0.1 M potassium phosphate, 1.25 mM EDTA, 0.01% Brij at pH 6.0,) for assaying.

Calpain I AssayCitation40

Calpain I from porcine erythrocytes (1.9 mg/mL) was purchased from Calbiochem. The incubation buffer was prepared from 2883 µL of HEPES pH 7.5 buffer, 39 µL of CaCl2 (0.5 M) and 78 µL of 0.5 M cysteine (freshly prepared). The incubation solution contained 30 µL inhibitor stock (10 µM–5 mM in DMSO), 30 µL enzyme stock (1.9 mg/mL) and 300 µL of incubation buffer at 25°C. Aliquots (50 µL) of the incubation solution were withdrawn at varying intervals and combined with 200 µL of substrate buffer (9524 µL HEPES pH 7.5 buffer, 80 µL 0.2 M Suc-Leu-Tyr-AMC in DMSO [Calbiochem Calpain I Substrate II], 214 µL 0.5 M cysteine and 107 µL 0.5 M CaCl2) for assaying.

Cruzain Assay 

The enzyme stock (50 µM) was diluted 2000 fold (25 nM) using assay buffer (100 mM sodium acetate pH 5.5, 5 mM DTT, 0.001% Triton X-100, freshly prepared). The incubation solution contained 12.5 µL inhibitor stock (10 µM-5 mM in DMSO), 12.5 µL D-2000 enzyme solution and 350 µL assay buffer. Aliquots (50 µL) of the incubation solution were withdrawn at varying intervals and combined with 200 µL of substrate buffer (4998 µL assay buffer and 2 µL 20 mMCbz-Phe-Arg-AMC in DMSO [Enzo Life Sciences]) for assaying.

Trypsin AssayCitation50

Trypsin from bovine pancreases was acquired from Sigma (Milwaukee, WI). The incubation solution contained 230 µL of buffer (50 mM sodium phosphate buffer, pH 8.0), 60 µL of trypsin (2.5 µM in buffer) and 10 µL inhibitor stock (10 µM-5 mM in DMSO) at 37°C. Aliquots (50 µL) of the incubation solution were withdrawn at varying intervals and combined with 200 µL of substrate buffer (4 µL of 22.7 mMCbz-Phe-Arg-AMC in DMSO [Enzo Life Sciences] in 2400 µL of buffer) for assaying.

Enzyme expression and purification

Cruzain

Cruzain was expressed and purified using a modified version (Lee G., Craik CS, unpublished results) of a previously published protocolCitation51, activated and purified as recently describedCitation32.

To prevent self-degradation, active cruzain was inhibited with S-methyl methanethiosulfonate (MMTS), a covalent reversible inhibitor. MMTS inhibited cruzain was further concentrated to 50 µM in 100 mM MES pH 5.8, 500 mMNaCl and stored at −80 °C.

Results and discussion

Formation of the allyl sulfone

One of the key challenges in synthesizing the inhibitors was finding the correct vinyl-allyl sulfone isomerization conditions. The proposed mechanism of allyl isomerization occurs by deprotonation of the α-proton forming a delocalized intermediate (). Subsequent protonation of the carbon adjacent to the sulfone furnishes allyl sulfones 1420Citation52. It was found that the choice of base, temperature and reaction duration were all significant factors affecting the isomerization. Previously reported methods of isomerization used KOtBuCitation52, DBUCitation53 or n-BuLi/t-BuOOHCitation40. The use of 1 eq. of n-BuLi and 4.5 eq. of t-BuOOH at −20°C resulted in nearly complete isomerization while the other bases, regardless of duration or temperature, produced a mixture of compounds which were not purified further. Hine et al.Citation52,Citation54 demonstrated that with a relatively low free energy barrier of isomerization, an equilibrium exists between the two isomers (vinyl and allyl) resulting in incomplete isomerization. Along with this equilibrium, we observed an additional side product, formed by elimination of the sulfonyl moiety (). Elimination occurs after formation of the allyl sulfone, when the adjacent nitrogen is deprotonated, eliminating the sulfonyl group and forming an extended conjugated system. Higher temperatures (>−20°C) or fewer equivalents of t-BuOOH favoured elimination. Elimination was only observed for the Hfe-based compoundsCitation7,Citation8,Citation11–13. We purified one analogue side product (Mu-Phe-Hfe-Elim) and added it to the selection of compounds to be tested for inhibitor potency with the various proteases.

Scheme 2.  Isomerization equilibrium and elimination mechanism.

Scheme 2.  Isomerization equilibrium and elimination mechanism.

We suspect that the acidic property of the t-BuOOH partially attenuates the strength of the n-BuLi, giving rise to a milder, buffered environment. This has not previously been reported in the literature. The mild conditions seem to favour the formation of the allyl sulfone while minimizing further elimination of the sulfonyl group. Addition of n-BuLi in absence of t-BuOOH exclusively results in complete elimination of the sulfonyl group, with no allyl sulfone intermediate observed. Chelation of Li+55 as well as the “syn-effect,” forming a homoaromatic system between Li+ and the delocalized anionic intermediateCitation53,Citation56, may be factors that explain the requirement of n-BuLi instead of DBU or KOtBu, assisting in the stabilization of the delocalized intermediate.

Reaction time was also a key factor in successful isomerization. Compounds 9, 10 and 13 (R-Phe-Phe-VS-Ph) with a Phe residue in the P1 position required less than two hours to isomerize while the remaining vinyl sulfones (7, 8, 11 and 12; R-Phe-Hfe-VS-Ph) with a Hfe residue in the P1 position required more than nine hours. This variation in reaction rate has been primarily attributed to sterics. The increased mobility introduced by the additional methylene of the Hfe residue, compared to the Phe side chain, allows the phenyl ring to impede access of the base to the α-proton, slowing the reaction rate and extending the required isomerization time.

Stereochemistry of allyl sulfones

Following the synthesis of allyl sulfones 14 and 15 (Mu-Phe-L/D-Hfe-AS-Ph), the compounds were assayed for inhibitory potency with cathepsin B and cruzain. Although the two compounds have identical 1D 1H NMR spectra (all spectroscopic data is provided in the Supporting Appendix), they differ in their ability to inhibit the enzymes by an order of magnitude, indicative of a significant difference in structure (see Enzyme Kinetics section, ). The only variation in structure that would be undetectable by 1H NMR is the stereochemistry of the allyl sulfone double bond. Because of the presence of only a single proton on the double bond, no distinctive J-coupling is produced with either E or Z configurations. This difference in stereochemistry arises during the isomerization process. With a different spatial orientation, the steric bulk of the P1 residue side chain can either favour E or Z configuration.

In an effort to deduce the stereochemistry of both analogues, we used Nuclear Overhauser Effect Spectroscopy (NOESY) experiments based on the difference in the spatial arrangement around the double bond, which should produce unique spatial magnetic coupling. The E and Z isomers of the Mu-Phe-Hfe-AS-Ph structure were modelled using molecular dynamics calculations by MOE to predict Nuclear Overhauser Effect (NOE) potentials. Jones et al.Citation57 have demonstrated that using energetically minimized models to predict NOE effects produced less than 4% error from the experimentally observed NOE. An MMFF94x potential was used to minimize the energetics of the model molecules using a solvation parameter based on distance with a relative dielectric constant of 4.801 to match the solvation of the samples in chloroform during NMR acquisition. Distances were measured between the alkene vinyl proton and the benzyl methylene of the Hfe side chain. Inter-proton distances of 4.01 and 2.22 Å were calculated for the E and Z isomers, respectively (). Based on this information, a weak NOE should be observed for the E-isomer (3.5–5 Å) while the Z-isomer should display a strong NOE (<2.5 Å)Citation58.

Figure 2.  Energy minimized models of the E and Z isomers of Mu-Phe-Hfe-AS-Ph.

Figure 2.  Energy minimized models of the E and Z isomers of Mu-Phe-Hfe-AS-Ph.

We performed 2D NOESY experiments on the P1 Hfe-based allyl sulfones (14, 15, 18 and 19). This method could not be applied to the P1 Phe-based compounds (16, 17 and 20). With only a single methylene group directly adjacent to the allyl sulfone double bond, long range J-coupling between the vinyl proton and benzyl methylene of the Phe made the NOE indistinguishable from the residual COSY cross peaks. However, with an additional carbon present between the vinyl proton and the benzyl methylene in the P1 Hfe-based compounds, long range J-coupling did not disrupt the resolution of the NOE.

We focused on the cross peak magnitude between the 2.71 and 4.64 ppm resonances, representing the spatial coupling between the protons of benzyl methylene of the Hfe residue and the vinyl proton, respectively. The difference in spectra is shown in (14 in blue and 15 in red). The individual peaks were integrated using equivalent contour levels and normalized against multiple residual COSY peaks present in each spectrum (). The allyl sulfones formed from the L-Hfe vinyl sulfones 14 and 18 display a 30% weaker NOE compared with their D-Hfe analogues 15 and 19. This allowed us to conclude that the allyl sulfones derived from the D-Hfe vinyl sulfone analogues exhibit Z configuration while those derived from the L-Hfe vinyl sulfone analogues possess E configuration.

Table 1.  Quantification of relative NOE magnitude and the deduced stereochemistry.

Figure 3.  NOESY comparison of 14 (blue) and 15 (red).

Figure 3.  NOESY comparison of 14 (blue) and 15 (red).

Enzyme kinetics

The synthesized dipeptidyl allyl sulfones 1420 were assayed with cathepsin B, cruzain, calpain I and trypsin using fluorogenic enzyme assays and the incubation method. Pseudo-first order kinetics were applied to calculate kobs/[I] values ()Citation59. The Cbz-containing compounds Citation16,Citation17 unfortunately were insoluble in the assays. All the remaining compounds Citation14,Citation15,Citation18–21 displayed time dependent, irreversible inhibition.

Table 2.  Inhibitor potencies for proteases.

All inhibitors Citation14,Citation15,Citation18–21 demonstrated specificity for cysteine proteases, with no inhibition observed for trypsin, a serine protease with similar substrate specificity as cruzain. The same fluorogenic substrate, Cbz-Phe-Arg-AMC, was used to assay both enzymes. Additionally, specificity for clan CA, family C1 was observed with negligible inhibition (<10 M−1s−1) of calpain I (family C2). Within clan CA, all of the allyl sulfones (14, 15, 18 and 20) showed approximately 10-fold selectivity for cruzain over cathepsin B except compound 19 (Z-Mp-Phe-Hfe-AS-Ph), which demonstrated no selectivity between the two enzymes (200 ± 40 M−1s−1 for cathepsin B and 130 ± 10 M−1s−1 for cruzain).

Comparing the potencies of inhibitors 14 (P3 = Mu) versus 18 (P3 = Mp), the P3 group exhibited almost negligible effect on potency for cathepsin B (420 ± 110 M−1s−1 for 14 versus 570 ± 20 M−1s−1 for 18). In contrast, the Mu substituent achieved a 3-fold increase in potency for cruzain compared with the Mp group (6080 ± 1390 M−1s−1 for 14 versus 2310 ± 230 M−1s−1 for 18).

The allyl sulfone double bond stereogenicity also contributed significantly to inhibitory potency. Examining E-allyl sulfones (14 and 18) and Z-allyl sulfones (15 and 19), the E stereochemistry resulted in an order of magnitude greater kobs/[I] values for all cysteine proteases tested. Comparing isomers 14 and 15, kobs/[I] values for cruzain were 6080 ± 1390 M−1s−1 and 260 ± 80 M−1s−1, respectively. A similar difference in potency was observed for cathepsin B. Diastereomers 18 and 19 displayed the same trend as well.

The P1 amino acid residue selection also played a key role in inhibitor potency. Although the stereochemistry of compound 20 (Mp-Phe-Phe-AS-Ph) could not be determined using the previously discussed NOE method (section 2.2), the kinetic data provides a clear indication that regardless of double bond stereogenicity, the Phe residue in the P1 position resulted in a decrease in potency. For cruzain, compound 20 had an inhibition rate constant of 45 ± 1 M−1s−1. With the assumption that compound 20 would follow the same isomerization mechanism as other allyl sulfones formed from vinyl sulfones with L-stereochemistry in their P1 residue, it should likely have E configuration. In that case, it is 100-fold less potent than its P1 Hfe analogue 18 (2310 ± 230 M−1s−1). Even if the above assumption is not valid and compound 20 adopted Z stereochemistry, the Z-Hfe analogue 19 still remains more potent (130 ± 10 M−1s−1). The same trends were observed with cathepsin B with an even greater decrease in potency with the presence of the P1 Phe residue. Further optimization of additional analogues with P1 Phe residues as well as investigation into absolute stereochemistry was abandoned.

The most potent of the synthesized inhibitors was compound 14 (E-Mu-Phe-Hfe-AS-Ph) with a kobs/[I] of 6080 ± 1390 M−1s−1 for cruzain. For comparative purposes, compound 11 (Mp-Phe-L-Hfe-VS-Ph, also K11777)Citation42 was assayed with cruzain (221000 ± 8000 M−1s−1) and displayed 10-fold selectivity over cathepsin B (39300 ± 1300 M−1s−1). Although none of the inhibitors were able to achieve the potency of K11777, equivalent selectivity was maintained.

The precise mechanism of inhibition of the allyl sulfone moiety remains unknown. We have previously proposed possible irreversible mechanisms ()Citation40 including a SN2 attack by the active site thiolate on the carbon adjacent to the sulfone, resulting in the elimination of the sulfone (a); a Michael addition to the double bond and the resulting elimination of the sulfone (b); and the direct elimination of the sulfone group, forming the enamide, which is then subject to a Michael addition via the thiolate on the terminal alkene carbon (c).

Scheme 3.  Potential mechanisms of inhibition.

Scheme 3.  Potential mechanisms of inhibition.

Compound 21, a side product in our synthesis, contains the same conjugated π-system as the proposed intermediate of mechanism c (). When compound 21 was assayed with the four enzymes, it showed limited inhibition of the three clan CA cysteine proteases (< 10 M−1s−1) and no inhibition of trypsin. We speculated that based on the much lower inhibition of 21 as compared to allyl sulfones14 and 15, mechanism c can be ruled out as the mechanism of enzyme inhibition. Crystallization of an allyl sulfone bound to cruzain or cathepsin B remains unsuccessful and future work will hopefully elucidate the mechanism.

Conclusion

The successful synthesis of a series of novel dipeptidyl allyl sulfones was achieved by base catalyzed isomerization of peptidyl vinyl sulfones. It was found that stereochemistry of the vinyl sulfone P1 residue controlled the resulting stereochemistry of the allyl sulfones during isomerization. Stereogenic conformations of the two isomers of the P1 Hfe allyl sulfones were resolved using a combination of 2D NOESY experiments and computational modelling. Kinetic assays of dipeptidyl allyl sulfones revealed inhibitor selectivity for family C1 of the clan CA cysteine proteases, with 10-fold intra-family specificity for cruzain over cathepsin B. The Mu substituent was found to be the optimal P3 group while the P1 Hfe residue was crucial in maintaining inhibitory potency. Stereogenic configuration of the allyl sulfone played the most substantial role in inhibitory potency. Overall, (E)-Mu-Phe-Hfe-AS-Ph14 was shown to be the most potent inhibitor. Partial insight into the inhibition mechanism was gained by analysis of the eliminated product of the Hfe-based allyl sulfones (enamide 21), ruling out a potential mechanism-based inhibitor pathway. Future work in co-crystallization of a bound allyl sulfone inhibitor to cruzain or cathepsin B will be necessary to extend our understanding of the allyl sulfone inhibition mechanism. In addition, further modification of the current design will be necessary to achieve or exceed the potency of K11777.

Acknowledgments

We would like to thank James McKerrow and his lab at the University of California − San Francisco for providing cruzain for enzyme inhibition analysis.

Declaration of interest

This material is based upon work supported by the National Science Foundation under Grant No. 0922775. The Howard Hughes Medical Institute provided additional research funding.

References

  • Nicholson DW, Thornberry NA. Caspases: killer proteases. Trends Biochem Sci 1997;22:299–306.
  • Otto HH, Schirmeister T. Cysteine Proteases and Their Inhibitors. Chem Rev 1997;97:133–171.
  • Hedstrom L. Serine protease mechanism and specificity. Chem Rev 2002;102:4501–4524.
  • Tong L. Viral proteases. Chem Rev 2002;102:4609–4626.
  • Barrett AJ, Rawlings ND, Woessner JF. Handbook of Proteolytic Enzymes. 2nd Edition. San Diego: Elsevier; 2004.
  • Lecaille F, Kaleta J, Broemme D. Human and Parasitic Papain-Like Cysteine Proteases: Their role in physiology and pathology and recent developments in inhibitor design. J ChemInform 2003;34:4459–4488.
  • Rawlings ND, Barrett AJ, Bateman A. MEROPS: the peptidase database. Nucleic Acids Res 2010;38:D227–D233.
  • Sorimachi H, Suzuki K. The structure of calpain. J Biochem 2001;129:653–664.
  • Johnson P. Calpains (intracellular calcium-activated cysteine proteinases): structure-activity relationships and involvement in normal and abnormal cellular metabolism. Int J Biochem 1990;22:811–822.
  • Chapman HA, Riese RJ, Shi GP. Emerging roles for cysteine proteases in human biology. Annu Rev Physiol 1997;59:63–88.
  • Mueller-Steiner S, Zhou Y, Arai H, Roberson ED, Sun B, Chen J et al. Antiamyloidogenic and neuroprotective functions of cathepsin B: implications for Alzheimer’s disease. Neuron 2006;51:703–714.
  • Joyce JA, Baruch A, Chehade K, Meyer-Morse N, Giraudo E, Tsai FY et al. Cathepsin cysteine proteases are effectors of invasive growth and angiogenesis during multistage tumorigenesis. Cancer Cell 2004;5:443–453.
  • Yan S, Sameni M, Sloane BF. Cathepsin B and human tumor progression. Biol Chem 1998;379:113–123.
  • Turk V, Turk B, Turk D. Lysosomal cysteine proteases: facts and opportunities. EMBO J 2001;20:4629–4633.
  • Turk B, Turk D, Turk V. Lysosomal cysteine proteases: more than scavengers. Biochim Biophys Acta 2000;1477:98–111.
  • Brömme D, Klaus JL, Okamoto K, Rasnick D, Palmer JT. Peptidyl vinyl sulphones: a new class of potent and selective cysteine protease inhibitors: S2P2 specificity of human cathepsin O2 in comparison with cathepsins S and L. Biochem J 1996;315 (Pt 1):85–89.
  • Marquis RW, Ru Y, LoCastro SM, Zeng J, Yamashita DS, Oh HJ et al. Azepanone-based inhibitors of human and rat cathepsin K. J Med Chem 2001;44:1380–1395.
  • Godat E, Chowdhury S, Lecaille F, Belghazi M, Purisima EO, Lalmanach G. Inhibition of a cathepsin L-like cysteine protease by a chimeric propeptide-derived inhibitor. Biochemistry 2005;44:10486–10493.
  • Reis FC, Costa TF, Sulea T, Mezzetti A, Scharfstein J, Brömme D et al. The propeptide of cruzipain–a potent selective inhibitor of the trypanosomal enzymes cruzipain and brucipain, and of the human enzyme cathepsin F. FEBS J 2007;274:1224–1234.
  • Woo J-T, Sigeizumi S, Yamaguchi K, Sugimoto K, Kobori T, Tsuji T et al. Peptidyl aldehyde derivatives as potent and selective inhibitors of cathepsin L. Bioorg Med Chem Lett 1995;5:1501–1504.
  • Ahmed NK, Martin LA, Watts LM, Palmer J, Thornburg L, Prior J et al. Peptidyl fluoromethyl ketones as inhibitors of cathepsin B. Implication for treatment of rheumatoid arthritis. Biochem Pharmacol 1992;44:1201–1207.
  • Turk D, Podobnik M, Popovic T, Katunuma N, Bode W, Huber R et al. Crystal structure of cathepsin B inhibited with CA030 at 2.0-A resolution: A basis for the design of specific epoxysuccinyl inhibitors. Biochemistry 1995;34:4791–4797.
  • Eakin AE, Mills AA, Harth G, McKerrow JH, Craik CS. The sequence, organization, and expression of the major cysteine protease (cruzain) from Trypanosoma cruzi. J Biol Chem 1992;267:7411–7420.
  • McKerrow JH, Sun E, Rosenthal PJ, Bouvier J. The proteases and pathogenicity of parasitic protozoa. Annu Rev Microbiol 1993;47:821–853.
  • Aparicio IM, Scharfstein J, Lima AP. A new cruzipain-mediated pathway of human cell invasion by Trypanosoma cruzi requires trypomastigote membranes. Infect Immun 2004;72:5892–5902.
  • Chagas disease (American trypanosomiasis): Fact sheet N340. WHO Media Centre 2010; Available at: http://www.who.int/mediacentre/factsheets/fs340/en/index.html.
  • Garcia S, Ramos CO, Senra JF, Vilas-Boas F, Rodrigues MM, Campos-de-Carvalho AC et al. Treatment with benznidazole during the chronic phase of experimental Chagas’ disease decreases cardiac alterations. Antimicrob Agents Chemother 2005;49:1521–1528.
  • Boiani M, Piacenza L, Hernández P, Boiani L, Cerecetto H, González M et al. Mode of action of nifurtimox and N-oxide-containing heterocycles against Trypanosoma cruzi: is oxidative stress involved? Biochem Pharmacol 2010;79:1736–1745.
  • Coura JR, Castro SLd. A critical review on chagas disease chemotherapy. Mem Inst Oswaldo Cruz 2002;97:3–24.
  • Castro JA, de Mecca MM, Bartel LC. Toxic side effects of drugs used to treat Chagas’ disease (American trypanosomiasis). Hum Exp Toxicol 2006;25:471–479.
  • Powers JC, Asgian JL, Ekici OD, James KE. Irreversible inhibitors of serine, cysteine, and threonine proteases. Chem Rev 2002;102:4639–4750.
  • Mott BT, Ferreira RS, Simeonov A, Jadhav A, Ang KK, Leister W et al. Identification and optimization of inhibitors of Trypanosomal cysteine proteases: cruzain, rhodesain, and TbCatB. J Med Chem 2010;53:52–60.
  • Li Z, Patil GS, Golubski ZE, Hori H, Tehrani K, Foreman JE et al. Peptide alpha-keto ester, alpha-keto amide, and alpha-keto acid inhibitors of calpains and other cysteine proteases. J Med Chem 1993;36:3472–3480.
  • Brak K, Doyle PS, McKerrow JH, Ellman JA. Identification of a new class of nonpeptidic inhibitors of cruzain. J Am Chem Soc 2008;130:6404–6410.
  • Asgian JL, James KE, Li ZZ, Carter W, Barrett AJ, Mikolajczyk J et al. Aza-peptide epoxides: a new class of inhibitors selective for clan CD cysteine proteases. J Med Chem 2002;45:4958–4960.
  • Du X, Guo C, Hansell E, Doyle PS, Caffrey CR, Holler TP et al. Synthesis and structure-activity relationship study of potent trypanocidal thio semicarbazone inhibitors of the trypanosomal cysteine protease cruzain. J Med Chem 2002;45:2695–2707.
  • dos Santos Filho JM, Leite AC, de Oliveira BG, Moreira DR, Lima MS, Soares MB et al. Design, synthesis and cruzain docking of 3-(4-substituted-aryl)-1,2,4-oxadiazole-N-acylhydrazones as anti-Trypanosoma cruzi agents. Bioorg Med Chem 2009;17:6682–6691.
  • Hernandes MZ, Rabello MM, Leite AC, Cardoso MV, Moreira DR, Brondani DJ et al. Studies toward the structural optimization of novel thiazolylhydrazone-based potent antitrypanosomal agents. Bioorg Med Chem 2010;18:7826–7835.
  • Palmer JT, Rasnick D, Klaus JL, Brömme D. Vinyl sulfones as mechanism-based cysteine protease inhibitors. J Med Chem 1995;38:3193–3196.
  • Götz MG, Caffrey CR, Hansell E, McKerrow JH, Powers JC. Peptidyl allyl sulfones: a new class of inhibitors for clan CA cysteine proteases. Bioorg Med Chem 2004;12:5203–5211.
  • Ettari R, Nizi E, Di Francesco ME, Dude MA, Pradel G, Vicík R et al. Development of peptidomimetics with a vinyl sulfone warhead as irreversible falcipain-2 inhibitors. J Med Chem 2008;51:988–996.
  • Engel JC, Doyle PS, Hsieh I, McKerrow JH. Cysteine protease inhibitors cure an experimental Trypanosoma cruzi infection. J Exp Med 1998;188:725–734.
  • Jacobsen W, Christians U, Benet LZ. In vitro evaluation of the disposition of A novel cysteine protease inhibitor. Drug Metab Dispos 2000;28:1343–1351.
  • Doyle PS, Zhou YM, Engel JC, McKerrow JH. A cysteine protease inhibitor cures Chagas’ disease in an immunodeficient-mouse model of infection. Antimicrob Agents Chemother 2007;51:3932–3939.
  • Barr SC, Warner KL, Kornreic BG, Piscitelli J, Wolfe A, Benet L et al. A cysteine protease inhibitor protects dogs from cardiac damage during infection by Trypanosoma cruzi. Antimicrob Agents Chemother 2005;49:5160–5161.
  • Kerr ID, Lee JH, Farady CJ, Marion R, Rickert M, Sajid M et al. Vinyl sulfones as antiparasitic agents and a structural basis for drug design. J Biol Chem 2009;284:25697–25703.
  • Wadsworth DH, Schupp OE, Seus EJ, Ford JA. The stereochemistry of the phosphonate modification of the Wittig reaction. J Org Chem 1965;30:680–685.
  • Maring CJ, Stoll VS, Zhao C, Sun M, Krueger AC, Stewart KD et al. Structure-based characterization and optimization of novel hydrophobic binding interactions in a series of pyrrolidine influenza neuraminidase inhibitors. J Med Chem 2005;48:3980–3990.
  • Bieth JG. Theoretical and practical aspects of proteinase inhibition kinetics. Meth Enzymol 1995;248:59–84.
  • Walker B, Elmore DT. The irreversible inhibition of urokinase, kidney-cell plasminogen activator, plasmin and beta-trypsin by 1-(N-6-amino-n-hexyl)carbamoylimidazole. Biochem J 1984;221:277–280.
  • Eakin AE, McGrath ME, McKerrow JH, Fletterick RJ, Craik CS. Production of crystallizable cruzain, the major cysteine protease from Trypanosoma cruzi. J Biol Chem 1993;268:6115–6118.
  • Hine J, Linden S-M, Wang A, Thiagarajan V. Structural effects on rates and equilibriums. 21. Double-bond-stabilizing abilities of dimethylamino, alkylsulfonyl, and acetyl substituents. J Org Chem 1980;45:2821–2825.
  • Inomata K, Hirata T, Suhara H, Kinoshita H, Kotake H, Senda AH. Stereochemistry of the Conversion of γ-substituted (E)-vinylsulfones to the corresponding allyl sulfones. Determination of the relative degree of “syn-effect”. Chem Lett 1988;17:2009–2012.
  • Hine J, Skoglund MJ. Effects of 57 substituents on the stabilities of carbon-carbon double bonds. J Org Chem 1982;47:4766–4770.
  • Belostotskii AM, Albeck A, Hassner A. Asymmetric induction by a remote chiral substituent − computationally determined stereodifferentiation in michael additions of α-lithiated allyl sulfones. Eur J Org Chem 2007;2007:4837–4844.
  • Hirata T, Sasada Y, Ohtani T, Asada T, Kinoshita H, Senda H et al. ChemInform abstract: “Syn-effect” in the conversion of (e)-vinylic sulfones to the corresponding allylic sulfones. J ChemInform 1992;23.
  • Jones CR, Butts CP, Harvey JN. Accuracy in determining interproton distances using Nuclear Overhauser Effect data from a flexible molecule. Beilstein J Org Chem 2011;7:145–150.
  • Neuhaus D, Williamson MP. The nuclear overhauser effect in structural and conformational analysis. 2nd Edition. Wiley-VCH; 2000
  • Schnell S, Mendoza C. The condition for pseudo-first-order kinetics in enzymatic reactions is independent of the initial enzyme concentration. Biophys Chem 2004;107:165–174.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.