Next Article in Journal
Constructing Interconnected Hollow Mesopore Sn-Si Mixed Oxide Microspheres by Aerosol-Assisted Alkali Treatment with Enhanced Catalytic Performance in Baeyer-Villiger Oxidation
Next Article in Special Issue
Recent Applications of Flavin-Dependent Monooxygenases in Biosynthesis, Pharmaceutical Development, and Environmental Science
Previous Article in Journal
Functionalized Coal Fly Ash Is an Efficient Catalyst for Synthesizing Furfural from Xylose at a Low Catalyst Load
Previous Article in Special Issue
Generating Novel and Soluble Class II Fructose-1,6-Bisphosphate Aldolase with ProteinGAN
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Synthesis of Pyrrole Disulfides Catalyzed by Lipase in Ethanol

1
Key Laboratory of Molecular Enzymology and Engineering of Ministry of Education, School of Life Sciences, Jilin University, Changchun 130023, China
2
Institute for Interdisciplinary Biomass Functional Materials Studies, Jilin Engineering Normal University, Changchun 130052, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(12), 1493; https://doi.org/10.3390/catal13121493
Submission received: 15 November 2023 / Revised: 4 December 2023 / Accepted: 5 December 2023 / Published: 6 December 2023
(This article belongs to the Special Issue State-of-the-Art Enzyme Engineering and Biocatalysis in China)

Abstract

:
Disulfides, as fundamental scaffolds, are widely present in peptides, natural products, and pharmaceutical molecules. However, traditional synthesis of disulfides often involves the utilization of toxic reagents or environmentally unfriendly reaction conditions. In this work, a green and efficient method was developed for synthesizing pyrrole disulfides using β-ketothioamides and ethyl cyanoacetate as substrates, with lipase serving as a catalyst. Under the optimal conditions (β-Ketothioamides (1 mmol), ethyl cyanoacetate (1 mmol), PPL (200 U), and EtOH (5 mL)), lipase leads to the formation of pyrrole disulfides in yields of up to 88% at 40 °C. The related mechanism is also speculated in this paper. This approach not only presents a new application of lipase in enzyme catalytic promiscuity, but also offers a significant advancement in the synthetic pathway for pyrrole disulfides and aligns with the current mainstream research direction of green chemistry, contributing to the further development of environmentally friendly biocatalytic processes.

Graphical Abstract

1. Introduction

β-Ketothioamides (KTAs) are versatile intermediates in organic synthesis possessing carbonyl and thioamide functional groups. These functionalities make them intriguing and valuable for various synthesis transformations. On the one hand, the carbonyl group in KTAs renders them susceptible to nucleophilic attack, allowing them to act as substrates for nucleophilic addition and leading to the formation of many heterocyclic compounds [1]. On the other hand, the thioamide functional group in KTAs undergoes isomerization to generate thiol [2], allowing KTAs to participate in synthesis. Multiple heterocycles, such as thiazoles [3,4], piperidines [5], pyran [6] and fused heterocycles [7,8], have been efficiently constructed. Such heterocycles form the core structures of numerous bioactive molecules and pharmaceutical formulations [9,10].
Pyrrole disulfides represent a fusion between pyrrole (a pentagonal heterocyclic organic compound) and disulfide functional groups and have relevance across various disciplines, such as organic synthesis [11], material science [12,13], and biochemistry [14,15]. The typical synthesis of pyrrole disulfides often involves the use of malodorous and toxic hydrogen sulfide gas or environmentally unfriendly reaction conditions. In 2020, Hussein’s group reported a method to synthesize an unreported pyrrole disulfide using Lawesson’s reagent and tetracyanide at 40 °C, achieving a yield of over 92% (Scheme 1a) [16]. In 2022, El-Remaily and co-workers synthesized pyrrole disulfides by reacting hydrogen sulfide with 2-(2-oxo-2-phenylethyl) malononitrile in ethanol (Scheme 1b) [17]. In the same year, Li et al. designed an environmentally friendly synthesis method using KTAs and ethyl cyanoacetates as substrates, sodium carbonate as the base (1.5 equiv.), and acetonitrile as the reaction solvent to generate the target product via an umpolung reaction (Scheme 1c) [18]. Despite these advancements, the synthesis routes for pyrrole disulfides are yet to be fully developed. Additional sustainable and environmentally benign strategies need to be devised.
Enzymes are large biological molecules capable of catalyzing chemical reactions within living organisms. Their presence and activity ensure the normal functioning of life. Enzymes display a high degree of specificity toward the reactions they catalyze, a characteristic termed as “specificity”. However, when studied outside their natural context, some enzymes can catalyze a variety of different chemical reactions or act on multiple substrates, such as hemoproteins [19,20], Baeyer–Villiger monooxygenase [21], transaminases [22] and laccase [23]. This phenomenon is referred to as “catalytic promiscuity”. The ability of enzymes to catalyze unnatural reactions is undeniably a substantial discovery for the advancement of organic catalytic synthesis. Lipases are a class of enzymes ubiquitously found in nature with a vast reservoir of resources and types, making them one of the most extensively studied hydrolases. Their broad substrate compatibility, high stability under extreme conditions, excellent enantioselectivity, and environmentally benign characteristics have garnered widespread attention from scholars. In addition to catalyzing natural reactions such as hydrolysis, lipases possess exceptional abilities to catalyze nonnative reactions [24,25,26]. Lipases have successfully catalyzed chemical reactions such as the Michael addition [27], Knoevenagel reaction [28], Aldol reaction [29], Mannich reaction [30], and oxidation reactions [31]. Our group conducted a number of studies on lipase-catalyzed unnatural reactions in various solvents, including organic solvents [32], ionic liquids [33], and supercritical carbon dioxide [34]. The results fully illustrated the feasibility and strong potential of lipase-catalyzed organic synthesis.
Compared with traditional chemical synthesis, the enzymatic strategy exhibits more potent catalytic capabilities and aligns significantly with the principles of green chemistry. Building upon the achievements from previous works and motivated by our interest in the high-performance capabilities of KTAs, our group investigated the lipase-catalyzed synthesis of pyrrole disulfides under mild conditions using KTAs (1a) and ethyl cyanoacetate (2) as substrates. (Scheme 1d). To the best of our knowledge, this study provides the first example of the biocatalytic synthesis of pyrrole disulfides.

2. Results and Discussion

2.1. Effect of Lipase Source

In the initial stages, we chose KTA (1a) and ethyl cyanoacetate (2) as template substrates to screen for the optimal reaction conditions. We also evaluated the catalytic abilities of lipases from various sources. As depicted in Table 1, all the selected lipases could afford the desired product 3a (entries 1–6). Among them, PPL demonstrated the most superior catalytic performance with a yield of 88%. The other lipases exhibited markedly different catalytic effects, which could be attributed to the variations in their protein structure. Control experiments were also conducted by employing inactivated PPLs and bovine serum albumin (entries 7–9) as catalysts, and no product formation was observed. Consistent results were also found when the catalyst was absent in the reaction (entry 10). These findings highlighted the crucial role of the active site of lipase in catalyzing this reaction. Compared with PPL, the use of sodium carbonate as the catalyst resulted in a moderate yield of the desired product 3a (entry 11). Compared to previous study, when PPL was employed as the catalyst, higher yields were achieved under lower temperature, undoubtedly demonstrating the catalytic potential of lipase [18]. This finding indicated that the enzymatic method is efficient for the synthesis of pyrrole disulfides.

2.2. Effect of Solvents

In enzymatic reactions, the reaction solvent and temperature are two crucial factors that significantly influence the outcome [35,36]. As shown in Table 2, the reaction solvent can influence the reaction extent by affecting the enzyme’s conformation and the substrates’ solubility. Compared with nonpolar solvents, polar solvents generally exhibit better catalytic effects. High yields could be obtained (entry 1–6) when polar solvents such as acetonitrile, ethanol, and N, N-dimethylformamide were utilized as reaction mediums. Conversely, solvents such as dichloromethane, toluene, and hexane resulted in yields lower than expected (entry 7–9). These observations demonstrated the importance of polar solvents in maintaining the catalytic efficiency of enzymes. However, even though water is a polar solvent, the yield of the desired product was relatively low due to the limited solubility of the substrates in water (entry 1). Therefore, water cannot be considered as the optimal solvent. Considering that ethanol aligns more closely with green chemistry principles than acetonitrile [37], we adopted ethanol as the optimal solvent in this reaction.

2.3. Effect of Temperature

The effect of temperature on this reaction was investigated. In general, increasing the temperature enhances the likelihood of collisions between substrate molecules and the enzyme. However, excessive temperature can lead to enzyme deactivation, thereby hampering the catalytic process [38]. Considering the thermostability and catalytic efficiency of lipase, we assessed its catalytic activity across a temperature range from 20 °C to 80 °C to understand the effect of temperature on this reaction (Figure 1). The results were consistent with our expectations. Below 40 °C, the catalytic efficiency increased with the temperature. The yield increased slightly with the increase in temperature from 40 °C to 60 °C. However, beyond 60 °C, the yield declined due to lipase inactivation. Based on the above results, we determined that the optimal reaction temperature is 40 °C.

2.4. Effect of Lipase Dosage

The dosage of the biocatalyst has a significant impact on the reaction. Considering the economy and efficiency of this enzymatic method, we investigated the effect of various amounts of PPL from 0 to 350 U (Figure 2). The change was most noticeable when the enzyme dosage was in the range of 0–150 U. When the enzyme dosage surpassed 150 U, the change rate of yield diminished with the increase in enzyme dosage. The highest yield was achieved at a dosage of 200 U; further increase in enzyme dosage did not result in any significant improvement in yield. Given the cost of lipase, the dosage of 200 U is sufficient for the reaction.

2.5. Substrate Scope

After determining the optimal reaction conditions, we investigated the substrate scope of the reaction. As depicted in Table 3, the reaction showed good compatibility with various substituents (R1) on the phenyl ring, regardless of whether they were electron-withdrawing (3b3d) or electron-donating (3e3g) groups. The corresponding pyrrole disulfides were obtained with satisfactory yields ranging from 70% to 90%. Similarly, favorable results were obtained with various substituents (R2) on the other phenyl ring (3h3n), and the corresponding yields ranged from 72% to 87%. Even when the phenyl ring was replaced with a methyl group, a yield of 75% was still achieved (3o). We further explored the substrates with substituents at the R1 and R2 positions (3p3t), and the results remained commendable. Whether the functional groups on R1 and R2 were electron-withdrawing or electron-donating, lipase-catalyzed product yields manifested varying degrees of superiority compared to those obtained through the chemical method [18]. These findings demonstrated the broad substrate applicability of PPL in catalyzing the synthesis of pyrrole disulfides, with satisfactory yields for products containing various types of functional group substitutions. To demonstrate the practical feasibility of the lipase-catalyzed synthesis of pyrrole disulfides from KTAs and ethyl cyanoacetates, we designed a corresponding decagram-scale reaction. In a 200 mL round-bottom flask containing 50 mL of ethanol, 2.55 g of substrate 1a and 1.13 g of ethyl cyanoacetate were added. PPL was employed as the catalytic agent (2000 U), and the reaction proceeded at 40 °C for 24 h. After the completion of the reaction was confirmed through TLC, purification was carried out, resulting in a product yield of 2.95 g corresponding to an 81% yield. This value is higher than that reported for chemical synthesis (70% yield). These findings motivated us to improve the catalytic performance of lipase by enzyme engineering and directed evolution techniques. Immobilization is an efficient strategy that leads to a significant enhancement in enzyme stability, catalytic efficiency, and recyclability. Currently, we are also investigating methods for immobilizing PPL to further enhance the efficiency of this enzymatic system and will continue to report our findings in due course [39,40].

2.6. Mechanistic Speculation

Based on our initial findings and previous literature, we proposed a possible mechanism for this enzymatic reaction (Scheme 2) [21,41,42]. First, a thioenol is formed from KTA 1a via tautomerization and then oxidized by the dissolved molecular oxygen to produce a dimeric intermediate I. Similarly, substrate 2 is deprotonated by lipase, generating an anion. This anion rapidly reacts with intermediate I, undergoing C-C bond modification to form intermediate II. The secondary amine in intermediate II undergoes N-cyclization with the cyano carbon, producing intermediate III. Intermediate III then undergoes isomerization, resulting in a fully substituted pyrrole IV. Finally, product 3a is formed after oxidation and dimerization.

3. Materials and Methods

3.1. Materials

Novozym 435 (a commercial immobilized C. antarctica lipase B, 15,000 U/g, U: one unit of the enzyme activity was defined as the amount of enzyme required to hydrolyze 1 μmol of ethyl cyanoacetate per minute at 30 °C) was purchased from Sigma-Aldrich China Co. (Beijing, China). PPL (porcine pancreas lipase, 5600 U/g), Cal-B (C. antarctica lipase B, 10,000 U/mL), and CSL (Candida sp. lipase, 6400 U/g) were purchased from Shanghai Yuan Ye Biological Technology Company (Shanghai, China), and MML (Mucor miehei lipase, 7300 U/g) was purchased from Sigma-Aldrich China Co. (Beijing, China). All the other chemical reagents were purchased from commercial suppliers (Bide Pharmatech, Aladdin, Energy Chemical, Beijing, China). HRMS were obtained on an Ultima Global spectrometer with an ESI source. NMR spectra were recorded on Bruker 400 MHz spectrometers (see Figure S1–S20 in the Supplementary Materials). Chemical shifts are in ppm with CDCl3 as the internal standard. NMR data are presented as follows: chemical shift (δ ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad), and coupling constant in Hertz (Hz), integration. The experiments were performed triplicate, and all data were obtained based on the average values.

3.2. General Procedure for Synthesis of 1

Compounds 1 were synthesized according to the procedure reported in the literature [43]. A mixture of NaH (10 mmol), acetophenone (10 mmol) and 1,4-dioxane (8 mL) was stirred at room temperature. Isothiocyanatobenzene (10 mmol) was added dropwise, and stirring was continued at room temperature for 2 h. The solids were collected using filtration and washed with 1,4-dioxane (10 mL). The solids were dissolved with water and then slowly neutralized via stirring with HCl. After filtration, the filter cake was dried. The obtained substance does not require further purification and can be directly utilized in the subsequent steps.

3.3. General Procedure for Lipase-Catalyzed Synthesis of 3

PPL (200 U) was added to a stirred solution of thioamides 1 (1 mmol) and ethyl cyanoacetate 2 (1 mmol) in ethanol (5 mL); the reaction mixture was stirred at 40 °C for 24 h. After completion of the reaction as monitored with TLC, the solvent was concentrated under vacuum and the residue was purified using flash column chromatography on silica gel with petroleum ether/ethyl acetate (3:1, v/v) as the eluent to yield the desired product 3.

3.4. Data of Products

1a 3-oxo-N,3-diphenylpropanethioamide
Isolated yield: 50%, yellow solid, mp 99–100 °C (lit. (25) 102–103 °C).
HRMS (ESI-TOF, [M + Na]+): calcd for C16H15NNaOS 292.0772, found 292.0768
1H NMR (400 MHz, CDCl3, δ, ppm) 10.93 (s, 1H), 7.08 (d, J = 7.6 Hz, 2H), 7.55–7.51 (m, 1H), 7.47–7.40 (m, 4H), 7.22–7.20 (m, 2H), 4.65 (s, 2H), 2.37 (s, 3H);
13C NMR (100 MHz, CDCl3, δ, ppm) 197.1, 191.0, 137.0, 134.5, 129.5, 129.0, 128.8, 128.6, 123.6, 54.0, 21.2;
3a diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 88% (321 mg), yellow solid, mp: 198–200 °C.
HRMS (ESI–TOF, [M + H]+): calcd for C40H35N4O6S2, 731.1993; found, 731.1992.
1H NMR (400 MHz, CDCl3) δ 7.98 (d, J = 7.6 Hz, 2H), 7.49 (t, J = 7.6 Hz, 1H), 7.42 (s, 3H), 7.33 (s, 1H), 7.25–7.16 (m, 2H), 5.19 (s, 2H), 3.97 (d, J = 7.2 Hz, 2H), 0.78 (t, J = 7.2 Hz, 3H).
3b diethyl 5,5′-disulfanediylbis (2-amino-4-(4-chlorobenzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 79% (316 mg), yellow solid, mp: 154–156 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Cl2, 821.1033; found, 821.1038.
1H NMR (400 MHz, CDCl3) δ 7.93 (d, J = 8.0 Hz, 2H), 7.44 (tt, J = 6.8, 3.2 Hz, 4H), 7.37–7.29 (m, 2H), 7.20 (s, 1H), 5.22 (s, 2H), 4.01 (q, J = 7.2 Hz, 2H), 0.85 (t, J = 7.2 Hz, 3H).
3c diethyl 5,5′-disulfanediylbis (2-amino-4-(4-bromobenzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 90% (399 mg), yellow solid, mp: 154–156 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Br2, 909.0028; found, 909.0037.
1H NMR (400 MHz, CDCl3) δ 7.86 (d, J = 8.0 Hz, 2H), 7.49 (d, J = 8.2 Hz, 5H), 7.20 (s, 1H), 5.22 (s, 2H), 4.02 (q, J = 7.2 Hz, 2H), 0.86 (t, J = 7.2 Hz, 3H).
3d diethyl 5,5′-disulfanediylbis(2-amino-4-(3-bromobenzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 75% (333 mg), yellow solid, mp: 180–182 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Br2, 909.0028; found, 909.0037.
1H NMR (400 MHz, CDCl3) δ 8.16 (s, 1H), 7.86 (d, J = 8.0 Hz, 1H), 7.71–7.50 (m, 1H), 7.45 (s, 3H), 7.26–7.13 (m, 2H), 5.25 (s, 2H), 4.00 (q, J = 7.2 Hz, 2H), 0.84 (t, J = 7.2 Hz, 3H).
3e diethyl 5,5′-disulfanediylbis (2-amino-4-(2-bromobenzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 70% (311 mg), yellow solid, mp: 180–182 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Br2, 909.0028; found, 909.0037.
1H NMR (400 MHz, CDCl3) δ 8.16 (s, 1H), 7.86 (d, J = 8.0 Hz, 1H), 7.45 (s, 3H), 7.18 (d, J = 8.4 Hz, 2H), 5.25 (s, 2H), 4.00 (q, J = 7.2 Hz, 2H), 0.84 (t, J = 7.2 Hz, 3H).
3f diethyl 5,5′-disulfanediylbis (2-amino-4-(4-methylbenzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 82% (311 mg), yellow solid, mp: 172–174 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H38N4O6NaS2, 781.2130; found, 781.2135.
1H NMR (400 MHz, CDCl3) δ 7.88 (d, J = 7.6 Hz, 2H), 7.41 (s, 3H), 7.16 (d, J = 23.8 Hz, 4H), 5.18 (s, 2H), 4.00 (q, J = 7.2 Hz, 2H), 2.40 (s, 3H), 0.83 (t, J = 7.2 Hz, 3H).
3g diethyl 5,5′-disulfanediylbis (2-amino-4-(4-(tert-butyl)benzoyl)-1-phenyl-1H-pyrrole-3-carboxylate)
Isolated yield: 71% (299 mg), yellow solid, mp: 188–190 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C48H50N4O6NaS2, 865.309; found, 865.314.
1H NMR (400 MHz, CDCl3) δ 7.92 (d, J = 8.0 Hz, 2H), 7.48–7.32 (m, 5H), 7.18 (s, 1H), 5.21 (s, 2H), 3.97 (d, J = 7.6 Hz, 2H), 1.37–1.23 (m, 9H), 0.72 (s, 3H).
3h diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(4-fluorophenyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 85% (326 mg), yellow solid, mp: 179–181 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6F2NaS2, 789.1629; found, 789.1637.
1H NMR (400 MHz, CDCl3) δ 7.96 (d, J = 7.6 Hz, 2H), 7.51 (t, J = 7.6 Hz, 1H), 7.38–7.28 (m, 1H), 7.25–7.07 (m, 4H), 5.16 (s, 2H), 3.95 (q, J = 7.2 Hz, 2H), 0.76 (t, J = 7.2 Hz, 3H).
3i diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(4-chlorophenyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 87% (348 mg), yellow solid, mp: 182–184 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Cl2, 821.1033; found, 821.1038.
1H NMR (400 MHz, CDCl3) δ 7.95 (d, J = 7.6 Hz, 2H), 7.51 (t, J = 7.2 Hz, 1H), 7.43–7.39 (m, 2H), 7.35 (d, J = 7.2 Hz, 2H), 7.26–7.09 (m, 2H), 5.19 (s, 2H), 3.95 (d, J = 7.2 Hz, 2H), 0.76 (d, J = 7.2 Hz, 3H).
3j diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(4-bromophenyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 84% (373 mg), yellow solid, mp: 238–240 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H32N4O6NaS2Br2, 909.0028; found, 909.0037.
1H NMR (400 MHz, CDCl3) δ 7.99–7.91 (m, 2H), 7.61–7.46 (m, 3H), 7.38–7.31 (m, 2H), 7.12 (s, 2H), 5.19 (s, 2H), 3.95 (q, J = 7.2 Hz, 2H), 0.85–0.72 (m, 3H).
3k diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(p-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 80% (303 mg), yellow solid, mp: 210–212 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H38N4O6NaS2, 781.2130; found, 781.2135.
1H NMR (400 MHz, CDCl3) δ 7.98 (d, J = 7.6 Hz, 2H), 7.55–7.42 (m, 1H), 7.34 (dd, J = 10.8, 3.6 Hz, 2H), 7.21 (d, J = 8.0 Hz, 2H), 7.07 (s, 2H), 5.17 (s, 2H), 3.96 (q, J = 7.2 Hz, 2H), 2.39 (s, 3H), 0.77 (t, J = 7.2 Hz, 2H).
3l diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(4-isopropylphenyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 76% (310 mg), yellow solid, mp: 220–222 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C46H46N4O6NaS2, 837.277; found, 837.282.
1H NMR (400 MHz, CDCl3) δ 7.81 (d, J = 7.6 Hz, 2H), 7.45 (t, J = 7.2 Hz, 1H), 7.29 (d, J = 12.4 Hz, 2H), 5.37 (d, J = 22.8 Hz,2H), 3.87 (d, J = 7.6 Hz, 2H), 1.45 (d, J = 6.8 Hz, 6H), 0.66 (t, J = 7.2 Hz, 3H).
3m diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(m-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 72% (273 mg), yellow solid, mp: 212–214 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H38N4O6NaS2, 781.2130; found, 781.2135.
1H NMR (400 MHz, CDCl3) δ 8.02–7.95 (m, 2H), 7.52–7.30 (m, 4H), 7.22 (d, J = 8.8 Hz, 1H), 6.99 (s, 2H), 5.20 (s, 2H), 3.97 (q, J = 7.2 Hz, 2H), 2.33 (s, 3H), 0.78 (t, J = 7.2 Hz, 3H).
3n diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-(o-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 77% (292 mg), yellow solid, mp: 210–212 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H38N4O6NaS2, 781.2130; found, 781.2135.
1H NMR (400 MHz, CDCl3) δ 7.79–7.72 (m, 2H), 7.57 (tt, J = 15.2, 7.2 Hz, 3H), 7.43–7.33 (m, 4H), 5.58 (s, 2H), 4.34 (q, J = 7.2 Hz, 2H), 2.21 (s, 3H), 0.85 (t, J = 7.2 Hz, 3H).
3o diethyl 5,5′-disulfanediylbis (2-amino-4-benzoyl-1-methyl-1H-pyrrole-3-carboxylate)
Isolated yield: 75% (227 mg), yellow solid, mp: 190–192 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C30H30N4O6NaS2, 629.1970; found, 629.2135.
1H NMR (400 MHz, CDCl3) δ 7.49 (d, J = 7.2 Hz, 3H), 7.35–7.29 (m, 3H), 5.10 (d, J = 13.6 Hz, 2H), 3.89 (dq, J = 9.2, 7.2 Hz, 2H), 3.33 (s, 3H), 0.71 (q, J = 7.6 Hz, 3H).
3p diethyl 5,5′-disulfanediylbis (2-amino-4-(4-methylbenzoyl)-1-(p-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 73% (287 mg), yellow solid, mp: 160–162 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C44H42N4O6NaS2, 809.2450; found, 809.2455.
1H NMR (400 MHz, CDCl3) δ 7.86 (d, J = 7.6 Hz, 2H), 7.19 (s, 6H), 5.15 (s, 2H), 3.98 (q, J = 7.2 Hz, 2H), 2.39 (d, J = 3.6 Hz, 6H), 0.82 (d, J = 7.2 Hz, 3H).
3q diethyl 5,5′-disulfanediylbis (2-amino-4-(4-chlorobenzoyl)-1-(p-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 77% (319 mg), yellow solid, mp: 134–136 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H36N4O6Cl2NaS2, 849.1351; found, 849.1345.
1H NMR (400 MHz, CDCl3) δ 8.07 (s, 1H), 7.58 (s, 1H), 7.46 (t, J = 10.4 Hz, 2H), 7.36 (t, J = 7.2 Hz, 2H), 6.90 (s, 2H), 5.03 (s, 2H), 3.95 (q, J = 7.2 Hz, 2H), 2.28 (s, 3H), 0.80 (t, J = 7.2 Hz, 3H).
3r diethyl 5,5′-disulfanediylbis (2-amino-1-(4-chlorophenyl)-4-(4-methylbenzoyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 80% (331 mg), yellow solid, mp: 134–136 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H36N4O6Cl2NaS2, 849.1351; found, 849.1345.
1H NMR (400 MHz, CDCl3) δ 7.86 (d, J = 7.6 Hz, 2H), 7.26–7.05 (m, 6H), 5.16 (s, 2H), 3.98 (q, J = 7.2 Hz, 2H), 2.41 (s, 3H), 0.81 (d, J = 8.4 Hz, 3H).
(3s) diethyl 5,5′-disulfanediylbis (2-amino-4-(4-bromobenzoyl)-1-(4-bromophenyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 81% (423 mg), yellow solid, mp: 188–190 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C40H30N4O6NaS2Br4, 1068.8299; found, 1068.8304.
1H NMR (400 MHz, CDCl3) δ 7.81 (d, J = 8.0 Hz, 2H), 7.48 (d, J = 8.0 Hz, 2H), 7.10 (s, 2H), 5.23 (s, 2H), 3.98 (dd, J = 17.2, 8.0 Hz, 2H), 0.84 (t, J = 7.2 Hz, 3H).
(3t) diethyl 5,5′-disulfanediylbis (2-amino-4-(4-bromobenzoyl)-1-(o-tolyl)-1H-pyrrole-3-carboxylate)
Isolated yield: 79% (362 mg), yellow solid, mp: 181–183 °C.
HRMS (ESI-TOF, [M + Na]+): calcd for C42H36N4O6NaS2Br2, 937.0044; found, 937.0049.
1H NMR (400 MHz, CDCl3) δ 7.84 (d, J = 8.0 Hz, 2H), 7.47 (d, J = 8.0 Hz, 2H), 7.31–7.20 (m, 2H), 6.98 (s, 2H), 5.23 (s, 2H), 4.01 (q, J = 7.2 Hz, 2H), 2.33 (s, 3H), 0.86 (t, J = 7.2 Hz, 3H).

4. Conclusions

We successfully developed a lipase-catalyzed method for the synthesis of pyrrole disulfides using KTAs and ethyl cyanoacetate as substrates. As the biocatalyst, PPL offers several advantages such as being environmentally friendly and achieving high yields. This enzymatic method utilized various substrates, and the transformation was completed in ethanol at 40 °C, with yields ranging from 70% to 90%. Moreover, it achieved a promising yield (70%) in a scale-up experiment. Simultaneously, based on control experiments, we identified the crucial role of the active center of lipase in catalyzing this reaction and speculated the reaction mechanism for the synthesis of pyrrole disulfides.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13121493/s1, Figure S1: 1H NMR of 3a. Figure S2. 1H NMR of 3b. Figure S3. 1H NMR of 3c. Figure S4. 1H NMR of 3d. Figure S5. 1H NMR of 3e. Figure S6. 1H NMR of 3f. Figure S7. 1H NMR of 3g. Figure S8. 1H NMR of 3h. Figure S9. 1H NMR of 3i. Figure S10. 1H NMR of 3j. Figure S11. 1H NMR of 3k. Figure S12. 1H NMR of 3l. Figure S13. 1H NMR of 3m. Figure S14. 1H NMR of 3n. Figure S15. 1H NMR of 3o. Figure S16. 1H NMR of 3p. Figure S17. 1H NMR of 3q. Figure S18. 1H NMR of 3r. Figure S19. 1H NMR of 3s. Figure S20. 1H NMR of 3t.

Author Contributions

Conceptualization, Z.W. and L.W.; methodology, F.W. and F.L.; formal analysis, F.W., Y.X., F.L. and J.M.; investigation, F.W.; writing—original draft, Y.X., Z.W. and L.W.; writing—review and editing, H.Z. and L.W.; visualization, F.W. and J.M.; supervision, H.Z. and L.W.; funding acquisition, Z.W. and L.W. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge the Science and Technology Development Program of Jilin Province (No. 20230101135JC).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pali, P.; Yadav, D.; Sahoo, S.C.; Shankar Singh, M. Metal-Free One-Pot Annulative Coupling of 2-Hydroxybenzaldehydes with β-Ketothioamides: Access to Diverse 2-Arylimino-2H-Chromenes. J. Org. Chem. 2022, 87, 12342–12351. [Google Scholar] [CrossRef]
  2. Zhang, L.-M.; Wen, L.-R.; Niu, X.-D.; Guo, W.-S.; Li, M. InCl3 catalyzed synthesis of thiophene derivatives via vinyl azides and β-ketothioamides. Tetrahedron 2023, 138, 133412. [Google Scholar] [CrossRef]
  3. Khan, S.; Ansari, M.A.; Singh, M.S. Access to Functionalized Thiazolothiadiazoles via the Chemoselective Cascade Heteroannulation of Thioamides with Hypervalent Iodine Reagents. Org. Lett. 2023, 25, 642–646. [Google Scholar] [CrossRef] [PubMed]
  4. Shukla, G.; Srivastava, A.; Singh, M.S. Metal- and Catalyst-Free, Formal [4 + 1] Annulation via Tandem C═O/C═S Functionalization: One-Pot Access to 3,5-Disubstituted/Annulated Isothiazoles. Org. Lett. 2016, 18, 2451–2454. [Google Scholar] [CrossRef]
  5. Lu, H.; Tan, C.-Y.; Zhang, H.-X.; Zhang, J.-L.; Liu, J.-Y.; Li, H.-Y.; Xu, P.-F. Participation of β-Ketothioamides in N-Heterocyclic Carbene-Catalyzed [3 + 3] Spiroannulation: Asymmetric Synthesis of Functionalized Spiro-piperidinone Derivatives. J. Org. Chem. 2018, 83, 15245–15255. [Google Scholar] [CrossRef]
  6. Zeng, Q.; Huang, X.; Liu, M.; Yu, Z.; Xiao, Y. Synthesis of Trifluoromethylated 4H-Pyran and 4H-Thiopyran via Divergent Reaction of β-CF3-1,3-Enynes with β-Ketothioamides. Org. Lett. 2022, 24, 8186–8191. [Google Scholar] [CrossRef] [PubMed]
  7. Man, N.-N.; Wang, J.-Q.; Zhang, L.-M.; Wen, L.-R.; Li, M. Chemo-, Regio-, and Stereoselective Construction of Core Skeleton of Furo[2,3-b]pyrrole via Multicomponent Bicyclization Reaction. J. Org. Chem. 2017, 82, 5566–5573. [Google Scholar] [CrossRef] [PubMed]
  8. Guo, W.-S.; Xin, X.; Zhao, K.-L.; Wen, L.-R.; Li, M. Acid/base-controlled chemodivergent synthesis of two differently functionalized tetrahydroimidazo[1,2-a]pyridines. RSC Adv. 2015, 5, 70429–70432. [Google Scholar] [CrossRef]
  9. Guo, W.-S.; Wen, L.-R.; Li, M. β-Ketothioamides: Efficient reagents in the synthesis of heterocycles. Org. Biomol. Chem. 2015, 13, 1942–1953. [Google Scholar] [CrossRef]
  10. Jagodziński, T.S. Thioamides as Useful Synthons in the Synthesis of Heterocycles. Chem. Rev. 2003, 103, 197–228. [Google Scholar] [CrossRef]
  11. Mulina, O.M.; Doronin, M.M.; He, L.-N.; Terent’ev, A.O. Disulfides as versatile starting reagents: Effective sulfonylation of alkenes with disulfides under electrochemical conditions. Org. Chem. Front. 2023, 10, 3559–3566. [Google Scholar] [CrossRef]
  12. Li, L.; Li, Z.; Yoshimura, A.; Sun, C.; Wang, T.; Chen, Y.; Chen, Z.; Littlejohn, A.; Xiang, Y.; Hundekar, P.; et al. Vanadium disulfide flakes with nanolayered titanium disulfide coating as cathode materials in lithium-ion batteries. Nat. Commun. 2019, 10, 1764. [Google Scholar] [CrossRef] [PubMed]
  13. Whittingham, M.S. Lithium titanium disulfide cathodes. Nat. Energy 2021, 6, 214. [Google Scholar] [CrossRef]
  14. Fass, D.; Thorpe, C. Chemistry and Enzymology of Disulfide Cross-Linking in Proteins. Chem. Rev. 2018, 118, 1169–1198. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, S.; Xu, X.-L.; Chiu, J.; Bowley, S.; Wu, Y.; Hogg, P.J.; Fang, C. Extracellular Protein Disulfide Isomerase Cleaves Allosteric Disulfides in Histidine-Rich Glycoprotein to Regulate Thrombus Formation. Blood 2020, 136 (Suppl. S1), 11–12. [Google Scholar] [CrossRef]
  16. Arafa, W.A.A.; Hussein, M.F. Design, Sonosynthesis, Quantum-Chemical Calculations, and Evaluation of New Mono- and Bis-pyridine Dicarbonitriles as Antiproliferative Agents. Chin. J. Chem. 2020, 38, 501–508. [Google Scholar] [CrossRef]
  17. Abdelhamid, A.A.; Salama, K.S.M.; Elsayed, A.M.; Gad, M.A.; Ali Ali El-Remaily, M.A.E.A. Synthesis and Toxicological Effect of Some New Pyrrole Derivatives as Prospective Insecticidal Agents against the Cotton Leafworm, Spodoptera littoralis (Boisduval). ACS Omega 2022, 7, 3990–4000. [Google Scholar] [CrossRef]
  18. Meng, X.; Guo, W.; Nan, G.; Li, M. Synthesis of pyrrole disulfides via umpolung of β-ketothioamides. Org. Biomol. Chem. 2022, 20, 7609–7612. [Google Scholar] [CrossRef]
  19. Dunham, N.P.; Arnold, F.H. Nature’s Machinery, Repurposed: Expanding the Repertoire of Iron-Dependent Oxygenases. ACS Catal. 2020, 10, 12239–12255. [Google Scholar] [CrossRef] [PubMed]
  20. Xie, H.; Li, F.; Xu, Y.; Wang, C.; Xu, Y.; Wu, J.; Li, Z.; Wang, Z.; Wang, L. Vitreoscilla hemoglobin: A natural carbene transfer catalyst for diastereo- and enantioselective synthesis of nitrile-substituted cyclopropanes. Green Chem. 2023, 25, 6853–6858. [Google Scholar] [CrossRef]
  21. Xu, J.; Peng, Y.; Wang, Z.; Hu, Y.; Fan, J.; Zheng, H.; Lin, X.; Wu, Q. Exploiting Cofactor Versatility to Convert a FAD-Dependent Baeyer–Villiger Monooxygenase into a Ketoreductase. Angew. Chem. Int. Ed. 2019, 58, 14499–14503. [Google Scholar] [CrossRef] [PubMed]
  22. López-Iglesias, M.; González-Martínez, D.; Gotor, V.; Busto, E.; Kroutil, W.; Gotor-Fernández, V. Biocatalytic Transamination for the Asymmetric Synthesis of Pyridylalkylamines. Structural and Activity Features in the Reactivity of Transaminases. ACS Catal. 2016, 6, 4003–4009. [Google Scholar] [CrossRef]
  23. Wang, Y.-F.; Xu, H.; Feng, L.; Shen, X.-F.; Wang, C.; Huo, X.-K.; Tian, X.-G.; Ning, J.; Zhang, B.-J.; Sun, C.-P.; et al. Oxidative coupling of coumarins catalyzed by laccase. Int. J. Biol. Macromol. 2019, 135, 1028–1033. [Google Scholar] [CrossRef]
  24. Busto, E.; Gotor-Fernández, V.; Gotor, V. Hydrolases: Catalytically promiscuous enzymes for non-conventional reactions in organic synthesis. Chem. Soc. Rev. 2010, 39, 4504–4523. [Google Scholar] [CrossRef]
  25. Ding, X.; Dong, C.-L.; Guan, Z.; He, Y.-H. Concurrent Asymmetric Reactions Combining Photocatalysis and Enzyme Catalysis: Direct Enantioselective Synthesis of 2,2-Disubstituted Indol-3-ones from 2-Arylindoles. Angew. Chem. Int. Ed. 2019, 58, 118–124. [Google Scholar] [CrossRef]
  26. Li, X.; Hu, X.; Qiao, Y.; Lu, T.; Bai, Y.; Xiong, J.; Li, X.; Gou, Q.; Ge, J. Enzyme-bimetallic hybrid catalyst for one-pot chemoenzymatic reactions. Chem. Eng. J. 2023, 452, 139356. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Zhao, Y.; Gao, X.; Jiang, W.; Li, Z.; Yao, Q.; Yang, F.; Wang, F.; Liu, J. Kinetic model of the enzymatic Michael addition for synthesis of mitomycin analogs catalyzed by immobilized lipase from T. laibacchii. Mol. Catal. 2019, 466, 146–156. [Google Scholar] [CrossRef]
  28. Evitt, A.S.; Bornscheuer, U.T. Lipase CAL-B does not catalyze a promiscuous decarboxylative aldol addition or Knoevenagel reaction. Green Chem. 2011, 13, 1141–1142. [Google Scholar] [CrossRef]
  29. Mustafa, A.; Ramadan, R.; Niikura, F.; Inayat, A.; Hafez, H. Highly selective synthesis of glyceryl monostearate via lipase catalyzed esterification of triple pressed stearic acid and glycerin. Sustain. Energy Technol. Assess. 2023, 57, 103200. [Google Scholar] [CrossRef]
  30. Wu, L.-L.; Xiang, Y.; Yang, D.-C.; Guan, Z.; He, Y.-H. Biocatalytic asymmetric Mannich reaction of ketimines using wheat germ lipase. Catal. Sci. Technol. 2016, 6, 3963–3970. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Zhao, Y.; Jiang, W.; Yao, Q.; Li, Z.; Gao, X.; Liu, T.; Yang, F.; Wang, F.; Liu, J. Lipase-Catalyzed Oxidation of Cyclohexanone To Form ε-Caprolactone and Kinetic Modeling. ACS Sustain. Chem. Eng. 2019, 7, 13294–13306. [Google Scholar] [CrossRef]
  32. Xu, Y.; Li, F.; Ma, J.; Li, J.; Xie, H.; Wang, C.; Chen, P.; Wang, L. Lipase-Catalyzed Phospha-Michael Addition Reactions under Mild Conditions. Molecules 2022, 27, 7798. [Google Scholar] [CrossRef] [PubMed]
  33. Li, F.; Xu, Y.; Wang, C.; Wang, C.; Xie, H.; Xu, Y.; Chen, P.; Wang, L. Efficient synthesis of substituted pyrazoles Via [3+2] cycloaddition catalyzed by lipase in ionic liquid. Process Biochem. 2023, 124, 253–258. [Google Scholar] [CrossRef]
  34. Li, F.; Wang, C.; Xu, Y.; Gao, X.; Xu, Y.; Xie, H.; Chen, P.; Wang, L. Lipase-Catalyzed Synthesis of Anthrone Functionalized Benzylic Amines via a Multicomponent Reaction in Supercritical Carbon Dioxide. ChemistrySelect 2022, 7, e202104517. [Google Scholar] [CrossRef]
  35. Klibanov, A.M. Improving enzymes by using them in organic solvents. Nature 2001, 409, 241–246. [Google Scholar] [CrossRef] [PubMed]
  36. Haki, G.D.; Rakshit, S.K. Developments in industrially important thermostable enzymes: A review. Bioresour. Technol. 2003, 89, 17–34. [Google Scholar] [CrossRef]
  37. Prat, D.; Wells, A.; Hayler, J.; Sneddon, H.; McElroy, C.R.; Abou-Shehada, S.; Dunn, P.J. CHEM21 selection guide of classical- and less classical-solvents. Green Chem. 2016, 18, 288–296. [Google Scholar] [CrossRef]
  38. Geronimo, I.; Denning, C.A.; Heidary, D.K.; Glazer, E.C.; Payne, C.M. Molecular Determinants of Substrate Affinity and Enzyme Activity of a Cytochrome P450(BM3) Variant. Biophys. J. 2018, 115, 1251–1263. [Google Scholar] [CrossRef]
  39. Suo, H.; Xu, L.; Xu, C.; Qiu, X.; Chen, H.; Huang, H.; Hu, Y. Graphene Oxide Nanosheets Shielding of Lipase Immobilized on Magnetic Composites for the Improvement of Enzyme Stability. ACS Sustain. Chem. Eng. 2019, 7, 4486–4494. [Google Scholar] [CrossRef]
  40. Bolivar, J.M.; Woodley, J.M.; Fernandez-Lafuente, R. Is enzyme immobilization a mature discipline? Some critical considerations to capitalize on the benefits of immobilization. Chem. Soc. Rev. 2022, 51, 6251–6290. [Google Scholar] [CrossRef]
  41. Feng, X.-W.; Li, C.; Wang, N.; Li, K.; Zhang, W.-W.; Wang, Z.; Yu, X.-Q. Lipase-catalysed decarboxylative aldol reaction and decarboxylative Knoevenagel reaction. Green Chem. 2009, 11, 1933–1936. [Google Scholar] [CrossRef]
  42. Tang, Y.; Wang, C.; Xie, H.; Xu, Y.; Wang, C.; Du, C.; Wang, Z.; Wang, L. Green Synthesis of Spirooxindoles via Lipase-Catalyzed One-Pot Tandem Reaction in Aqueous Media. Catalysts 2023, 13, 143. [Google Scholar] [CrossRef]
  43. Feng, X.; Wang, J.-J.; Xun, Z.; Huang, Z.-B.; Shi, D.-Q. Multicomponent Strategy to Indeno[2,1-c]pyridine and Hydroisoquinoline Derivatives through Cleavage of Carbon–Carbon Bond. J. Org. Chem. 2015, 80, 1025–1033. [Google Scholar] [CrossRef]
Scheme 1. Previous works and lipase-catalyzed synthesis of pyrrole disulfides (Reprinted/adapted with permission from Refs. [16,17,18]). (a): Hussein’s work, Modather F. Hussein, 2020 [16], (b): El-Remaily’s work, Mahmoud Abd El Aleem Ali Ali El-Remaily [17], 2022, (c): Li’s work, Ming Li, 2022 [18]).
Scheme 1. Previous works and lipase-catalyzed synthesis of pyrrole disulfides (Reprinted/adapted with permission from Refs. [16,17,18]). (a): Hussein’s work, Modather F. Hussein, 2020 [16], (b): El-Remaily’s work, Mahmoud Abd El Aleem Ali Ali El-Remaily [17], 2022, (c): Li’s work, Ming Li, 2022 [18]).
Catalysts 13 01493 sch001
Figure 1. Effect of temperature on the enzymatic synthesis of pyrrole disulfide (3a). (Reaction conditions: 1a (1 mmol), 2 (1 mmol), EtOH (5 mL), PPL (200 U), 24 h).
Figure 1. Effect of temperature on the enzymatic synthesis of pyrrole disulfide (3a). (Reaction conditions: 1a (1 mmol), 2 (1 mmol), EtOH (5 mL), PPL (200 U), 24 h).
Catalysts 13 01493 g001
Figure 2. Effect of enzyme dosage on the enzymatic synthesis of pyrrole disulfide (3a). (Reaction conditions: 1a (1 mmol), 2 (1 mmol), EtOH (5 mL), PPL, 40 °C, 24 h).
Figure 2. Effect of enzyme dosage on the enzymatic synthesis of pyrrole disulfide (3a). (Reaction conditions: 1a (1 mmol), 2 (1 mmol), EtOH (5 mL), PPL, 40 °C, 24 h).
Catalysts 13 01493 g002
Scheme 2. Plausible mechanism for the enzymatic synthesis of pyrrole disulfides.
Scheme 2. Plausible mechanism for the enzymatic synthesis of pyrrole disulfides.
Catalysts 13 01493 sch002
Table 1. Lipase-catalyzed synthesis of pyrrole disulfide (3a) a.
Table 1. Lipase-catalyzed synthesis of pyrrole disulfide (3a) a.
Catalysts 13 01493 i001
EntryEnzyme bYield (%)
1PPL88
2Cal-B68
3Novozym 43577
4CSL40
5CRL48
6MML55
7BSA cN.D. d
8PPL eN.D.
9PPL fN.D.
10NoneN.D.
11Na2CO3 g80
a Reaction conditions: 1a (1 mmol), 2 (1 mmol), EtOH (5 mL), lipase (200 U), 40 °C, 24 h. b PPL (Porcine pancreatic lipase); Cal-B (C. antarctica lipase B); Novozym 435 (a commercial immobilized Cal-B); CSL (Candida sp. lipase); CRL (C. rugosa lipase); MML (Mucor miehei Lipase). c BSA (Bovine serum albumin). d No detected. e PPL was denatured by heating it to 100 °C for 24 h in water. f PPL was denatured by treating PPL with phenylmethanesulfonyl fluoride (PMSF). g Na2CO3 (1.5 mmol).
Table 2. Effect of solvents on the enzymatic synthesis of pyrrole disulfide (3a) a.
Table 2. Effect of solvents on the enzymatic synthesis of pyrrole disulfide (3a) a.
EntrySolventYield (%)
1Water73
2N, N-dimethylformamide68
3Dimethyl sulfoxide57
4Ethanol88
5Acetonitrile89
6Ethyl Acetate80
7Dichloromethane37
8Toluene36
9n-Hexane27
a Reaction conditions: 1a (1 mmol), 2 (1 mmol), solvent (5 mL), PPL (200 U), 40 °C, 24 h.
Table 3. Synthesis of pyrrole disulfides 3 catalyzed by lipase.
Table 3. Synthesis of pyrrole disulfides 3 catalyzed by lipase.
Catalysts 13 01493 i002
Catalysts 13 01493 i003
Reaction conditions: 1 (1 mmol), 2 (1 mmol), EtOH (5 mL), PPL (200 U), 40 °C, 24 h. In compound 3, the black and blue colors represent the source of the atoms (black from 1, blue from 2), while the red color indicates newly formed chemical bonds.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wen, F.; Xu, Y.; Li, F.; Ma, J.; Wang, Z.; Zhang, H.; Wang, L. Efficient Synthesis of Pyrrole Disulfides Catalyzed by Lipase in Ethanol. Catalysts 2023, 13, 1493. https://doi.org/10.3390/catal13121493

AMA Style

Wen F, Xu Y, Li F, Ma J, Wang Z, Zhang H, Wang L. Efficient Synthesis of Pyrrole Disulfides Catalyzed by Lipase in Ethanol. Catalysts. 2023; 13(12):1493. https://doi.org/10.3390/catal13121493

Chicago/Turabian Style

Wen, Feiyang, Yuelin Xu, Fengxi Li, Jinglin Ma, Zhi Wang, Hong Zhang, and Lei Wang. 2023. "Efficient Synthesis of Pyrrole Disulfides Catalyzed by Lipase in Ethanol" Catalysts 13, no. 12: 1493. https://doi.org/10.3390/catal13121493

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop