Next Article in Journal
Essential Oils and Their Compounds as Potential Anti-Influenza Agents
Next Article in Special Issue
Magnetic Polyethyleneimine Nanoparticles Fabricated via Ionic Liquid as Bridging Agents for Laccase Immobilization and Its Application in Phenolic Pollutants Removal
Previous Article in Journal
How Well Do We Handle the Sample Preparation, FT-ICR Mass Spectrometry Analysis, and Data Treatment of Atmospheric Waters?
Previous Article in Special Issue
Efficient Synthesis of Key Chiral Intermediate in Painkillers (R)-1-[3,5-Bis(trifluoromethyl)phenyl]ethanamine by Bienzyme Cascade System with R-ω-Transaminase and Alcohol Dehydrogenase Functions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lipase-Catalyzed Phospha-Michael Addition Reactions under Mild Conditions

1
Key Laboratory of Molecular Enzymology and Engineering of Ministry of Education, School of Life Sciences, Jilin University, Changchun 130023, China
2
State Key Laboratory of Supramolecular Structure and Materials, Jilin University, Changchun, 130023, China
3
The Second Hospital of Jilin University Changchun, Jilin University, Changchun 130041, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(22), 7798; https://doi.org/10.3390/molecules27227798
Submission received: 11 September 2022 / Revised: 4 November 2022 / Accepted: 9 November 2022 / Published: 12 November 2022

Abstract

:
Organophosphorus compounds are the core structure of many active natural products. The synthesis of these compounds is generally achieved by metal catalysis requiring specifically functionalized substrates or harsh conditions. Herein, we disclose the phospha-Michael addition reaction of biphenyphosphine oxide with various substituted β-nitrostyrenes or benzylidene malononitriles. This biocatalytic strategy provides a direct route for the synthesis of C-P bonds with good functional group compatibility and simple and practical operation. Under the optimal conditions (styrene (0.5 mmol), biphenyphosphine oxide (0.5 mmol), Novozym 435 (300 U), and EtOH (1 mL)), lipase leads to the formation of organophosphorus compounds in yields up to 94% at room temperature. Furthermore, we confirm the role of the catalytic triad of lipase in this phospha-Michael addition reaction. This new biocatalytic system will have broad applications in organic synthesis.

1. Introduction

The incorporation of phosphorus can beneficially modify the physicochemical and biological properties of bioactive molecules. Organophosphorus compounds are increasingly utilized in medicine and biological systems because phosphorus substituents can improve the pharmacokinetic and pharmacological properties of drugs, including their potency, cell permeability, and metabolic stability [1,2,3,4,5]. The phospha-Michael addition of phosphorus nucleophiles is one of the most practical approaches to generating new C-P bonds and transforming them into corresponding 1,4-addition products. [6,7,8,9,10,11,12].
Over the past decades, a general method for the conjugate transfer of phosphinyl groups from silylphosphines to α-β-unsaturated carbonyl catalyzed by a bench-stable palladium (II) catalyst has been reported (Scheme 1a) [13,14]. Song and co-workers reported the protocol for hydrophosphination of α-β-unsaturated compounds with diphe-nylphosphine catalyzed by nickel or palladium (Scheme 1b) [15,16,17,18,19]. Organic catalysts can mediate the formation of the phosphorus–carbon bond between biphenylphosphine oxide and the Michael acceptor, as reported by Bharati’s group (Scheme 1c); these guanidine derivative catalysts exhibit better catalytic ability than triethylamine [20,21,22]. Despite this progress, inconvenient and non-atom-economic substrates (such as silylphosphines), expensive metal catalysts, and hazardous organic solvents limit the practical application of the methods. Therefore, the development of new phospha-Michael addition reactions under mild conditions is still an urgent issue.
Enzymes, which catalyze metabolism in organisms, have the advantages of high catalytic efficiency and excellent selectivity and are more environmentally attractive than traditional chemical catalysts. In general, the most impressive enzymes are thought to only catalyze specific substrates from a natural selection; this property is called specificity, which limits the application of enzymes in industrial production. By contrast, “catalytic promiscuity” occurs when the catalytic abilities of the active site are used to catalyze a distinctly different type of reaction. Currently, hydrolase, hemoproteins, and artificial metalloenzymes have emerged as promising catalysts for non-natural reactions forming C-C and C-hetero bonds [23,24,25,26,27,28,29,30]. Lipases (EC 3.1.1.3) are the ideal biocatalyst due to their commercial availability, broad substrate specificity, and good stability in nonaqueous media. As one of the interfacial enzymes, the interfacial activation of lipases results primarily from conformational changes in the enzymes that expose the active site and provide a hydrophobic surface for interaction with the substrate [31]. These properties make lipases even more versatile in the rapidly developing field of catalytic promiscuity [32,33]. Our group and other researchers have previously reported that lipases can promote numerous promiscuous reactions, such as the Michael addition, aldol reaction, and oxidation reactions [34,35,36,37,38,39,40,41]. Compared with previous chemical methods, lipase-catalyzed non-natural reactions have many advantages, such as mild reaction conditions, environmental friendliness, and a green, reusable biocatalyst. As a continuation of our recent research exploring new non-natural reactions of lipase in organic synthesis, we have studied the reaction of dimethylphosphite with β-nitrostyrenes or benzylidenemalononitrile. In this work, we establish a convenient and efficient synthesis of organophosphorus compounds via a lipase-catalyzed phospha-Michael addition (Scheme 1d).

2. Results and Discussion

Initially, we selected β-nitrostyrene (1a) and biphenyphosphine oxide (2) as the model substrates for the lipase-catalyzed phospha-Michael addition reaction to optimize the conditions (Table 1). All of the selected lipases can catalyze the phospha-Michael addition reaction (entries 1–4). Novozym 435 (a commercial immobilized Cal-B) was the best at generating a yield of the product in 2 h (3a, entry 1), and the other lipases demonstrated relatively satisfactory yields in 4 h (entries 2–4). The above results are in accordance with the previous works about lipase-catalyzed Michael additions [42,43]. The catalytic ability of these lipases was higher than that of chemical catalysts (triethylamine, entry 5). Phenylmethanesulfonyl fluoride (PMSF) is usually used as an irreversible inhibitor of lipase for mechanism studies [44]. In this phospha-Michael addition reaction, Novozym 435 was inactivated by PMSF to confirm the role of the active center of the lipase, and the reaction did not proceed when the inactive lipase was used as the catalyst (entry 7). Similar results were observed in the thermally denatured lipase-catalyzed reaction and the control reaction (entries 8, 9). This result coincides with the lipase-catalyzed aza-Michael addition previously reported by us [45].
An appropriate reaction medium can notably improve the yields of enzymatic reactions [46]. Thus, various solvents were evaluated for the lipase-catalyzed Michael addition (Table 1, entries 1, 9–16). Novozym 435 showed good catalytic activity in polar solvents, such as MeCN, EtOH, and MeOH, but poor performance in dichloromethane, toluene, and hexane. However, the reaction did not show a satisfactory yield due to the poor solubility of the substrate in water (entry 14). Based on these results, polar solvents may help maintain the catalytic performance of enzymes. Novozym 435 showed the highest yield in EtOH (entry 1). Thus, EtOH was selected as the solvent in this reaction. Meanwhile, using green and low-cost media makes this method more environmentally friendly and economical than other synthesis processes for this reaction [47]. We then monitored the enantioselectivity for this phospha-Michael addition reaction. Product 3a was analyzed by chiral HPLC analysis. Unfortunately, neither triethylamine nor lipase can provide chiral induction in the reaction.
We also evaluated the effect of the enzyme dosage (Table 1, entries 1, 17, and 18). When the catalyst dosage varied from 150 U to 300 U, the yield increased with the increasing catalyst dosage. However, a higher enzyme dosage (>300 U) did not further increase the reaction yield. Therefore, from an economical point of view, 300 U was chosen as the optimum enzyme dosage. Meanwhile, the time course of the lipase-catalyzed phospha-Michael addition reaction was investigated in the supporting information (Figure S1).
Enzyme immobilization has been found to be a solution for many enzyme limitations: it can improve enzyme stability; alter enzyme selectivity, specificity, and activity; reduce inhibitions; enlarge the operation window and resistance to chemicals; and even be coupled with enzyme purification [48,49,50,51,52,53,54,55,56,57]. In this work, the reusability of Novozym 435 was investigated; the Novozym 435 was able to be recovered by filtration easily, washed with dichloromethane, and then reused directly. Figure 1 shows that as the number of cycles increases, the yield of 3a decreases; it obtained a 79% yield even after seven cycles. The decrease in the yield of 3a might be due to the leakage of proteins from the support and the slight dissolution of the support during the reaction cycles (Table S1).
Under the optimum reaction conditions for lipase-catalyzed phospha-Michael addition reactions (Table 1, entry1), various β-nitrostyrene derivatives (1) and biphenyphosphine oxide (2) were utilized to evaluate the scope of this enzymatic method (Table 2). When electron-donating substituents, such as methyl and methoxy groups, were attached to the phenyl ring, products 3b–3d and 3j were obtained in high yields. When the substituents on the phenyl ring were electron-withdrawing, such as Cl– and Br–, products 3e–3i were also obtained in satisfactory yields. Operation of this method is simple and practical, enabling the precipitation of the product isolated by filtration and washed with cold acetonitrile to obtain 3. The method also has excellent yields and diverse libraries of target compounds with potential biological activities.
Inspired by the above experiment, further examination of the substrate scope of the lipase-catalyzed phospha-Michael addition reaction was performed. Biphenyphosphine oxide (2) and benzylidene malononitrile (4) were chosen to evaluate the scope of this protocol (Table 3). The products have high yields regardless of whether the electron-withdrawing substituents or electron-donating substituents are attached to the phenyl of 4 (5a–5i). For the other aromatic rings, products 5j and 5k were also generated in good yields. These results demonstrate the broad substrate applicability of the lipase-catalyzed phospha-Michael reaction.
A possible reaction pathway for the lipase-catalyzed phospha-Michael addition reaction was speculated based on our preliminary results and previous reports (Scheme 2). First, 2a is deprotonated by the catalytic triad of lipase. Subsequently, a nucleophilic attack on nitroalkene can generate a zwitterionic intermediate I with high reactivity, which might be stabilized by the oxyanion of lipase. Finally, the protonation of intermediate I in the presence of the Asp-His dyad can provide the desired product 3a accompanied by the regeneration of the reactive lipase active site.

3. Materials and Methods

3.1. General Information

PPL (porcine pancreas lipase, 5600 U/g), PSL (Pseudomonas sp. lipase, 8500 U/g), CalB (Lipase B from Candida antarctica, 10,000 U/mL), and CSL (Candida sp. lipase, 6400 U/g) were purchased from Shanghai Yuan Ye Biological Technology Company and Novozym 435 (15,000 U/g) was purchased from Sigma-Aldrich China Co. (Beijing, China). One unit of the enzyme activity was defined as the amount of enzyme required to hydrolyze 1 μmol of p-nitrophenyl acetate per minute at 30 °C. All other chemical reagents were purchased from commercial suppliers (Bide Pharmatech, Aladdin, Energy Chemical). All of the commercially available reagents and solvents were used without further purification. Proton nuclear magnetic resonance (1H NMR) spectra were recorded on a 400 MHz spectrometer in CDCl3 or DMSO. Chemical shifts of protons are reported in parts per million downfield from tetramethyl silane (TMS) and are referenced to residual protium in the NMR solvent (CHCl3 = δ 7.26 ppm, DMSO = δ 2.50 ppm). NMR data are presented as follows: chemical shift (δ ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad), coupling constant in Hertz (Hz), integration. The experiments were performed in triplicate, and all data were obtained based on the average values.

3.2. General Procedure for the Synthesis of 3

To create a mixture of β-nitrostyrene (1, 0.5 mmol) and biphenyphosphine oxide (2, 0.5 mmol) in EtOH (1 mL), lipase (protein content: 300 U) was added and stirred in a room temperature shaker until completion of the reaction as indicated by TLC. The resulting precipitate was then filtered and washed with cold ethanol to obtain the compounds of 3 as pure products requiring no further purification. The products were further characterized by NMR spectroscopy.

3.3. General Procedure for the Synthesis of 5

To create a mixture of benzylidene malononitrile (4, 0.5 mmol) and biphenyphosphine oxide (2, 0.5 mmol) in EtOH (1 mL), lipase (protein content: 300 U) was added and stirred in a room temperature shaker until completion of the reaction as indicated by TLC. After completion of the reaction, the reaction mass was extracted with ethyl acetate (3 × 5 mL). The organic layers were combined and subjected to drying by a rota-evaporator to obtain crude 5, which was purified using column chromatography (hexane-EtOAc).

3.4. Synthetic Procedures of Benzylidene Malono-Nitrile 4

Benzaldehyde (10 mmol), malononitrile (9 mmol), and DABCO (0.1 mmol) were added to a round-bottom flask with a 25 mL capacity, followed by the addition of 10 mL of water, and stirred in a room temperature shaker for 12 h. Upon completion of the reaction, as indicated by TLC, a precipitate came out from the mixture. The precipitate was collected by filtration, washed with n-hexane several times, and dried to obtain benzylidene malononitrile (4).

3.5. The Denaturation of Novozym 435 Treated by PMSF

Novozym 435 (20 mg/mL) was incubated in ethanol with the PMSF (phenylmethanesulfonyl fluoride) stock solution (10 mmol/L in ethanol) at room temperature to denature the lipase. The reaction system was filtered, and the residue was washed with ethanol to remove the unbound PMSF. Then, the denatured Novozym 435 was dried overnight under a vacuum.

3.5.1. 3a ((2-nitro-1-phenylethyl) diphenylphosphine oxide)

1H NMR (400 MHz, DMSO-d6) δ 8.18–8.10 (m, 2H), 7.78–7.70 (m, 2H), 7.68–7.62 (m, 3H), 7.41 (td, J = 7.2, 1.6 Hz, 3H), 7.37–7.31 (m, 2H), 7.23–7.12 (m, 3H), 5.15–4.97 (m, 2H), 4.91–4.81 (m, 1H). 13C NMR (151 MHz, DMSO) δ 133.20, 132.62, 131.52, 131.21, 131.15, 131.06, 131.00, 130.62, 130.41, 129.81, 129.73, 128.98, 128.90, 128.83, 128.37, 76.47, 76.06, 44.44, 44.02. 31P NMR (243 MHz, DMSO) δ 30.84.

3.5.2. 3b ((2-nitro-1-(p-tolyl) ethyl) diphenylphosphine oxide)

1H NMR (400 MHz, Chloroform-d) δ 8.01–7.94 (m, 2H), 7.67–7.56 (m, 3H), 7.51–7.45 (m, 2H), 7.40 (dd, J = 7.2, 1.6 Hz, 1H), 7.30 (td, J = 6.4, 5.6, 3.6 Hz, 2H), 7.17 (dd, J = 8.0, 2.0 Hz, 2H), 7.02 (d, J = 7.6 Hz, 2H), 5.11–5.01 (m, 1H), 4.77–4.71 (m, 1H), 4.41 (s, 1H), 2.26 (d, J = 1.6 Hz, 3H). 13C NMR (151 MHz, CDCl3) δ 135.71, 130.35, 129.74, 128.85, 128.79, 128.69, 128.23, 128.09, 127.59, 127.40, 127.15, 126.91, 126.05, 125.97, 114.09, 74.98, 74.77, 74.56, 73.57, 73.53, 43.28, 42.86, 18.73. 31P NMR (243 MHz, CDCl3) δ 24.87.

3.5.3. 3c ((1-(4-methoxyphenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, DMSO-d6) δ 8.15–8.07 (m, 2H), 7.81–7.72 (m, 2H), 7.65 (dt, J = 6.0, 3.6 Hz, 3H), 7.45–7.32 (m, 5H), 6.80–6.74 (m, 2H), 5.06–4.90 (m, 2H), 4.84–4.77 (m, 1H), 3.67 (s, 3H). 13C NMR (151 MHz, DMSO) δ 159.06, 133.08, 132.53, 131.18, 131.07, 130.28, 129.75, 129.67, 128.99, 128.91, 124.34, 114.20, 76.34, 55.38, 55.06, 43.58, 43.15. 31P NMR (243 MHz, DMSO) δ 30.56.

3.5.4. 3d ((1-(4-isopropylphenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, Chloroform-d) δ 8.01 (t, J = 9.2 Hz, 2H), 7.66 (dq, J = 14.0, 7.2 Hz, 3H), 7.43 (q, J = 7.6, 6.4 Hz, 3H), 7.31 (d, J = 2.8 Hz, 1H), 7.28 (s, 1H), 7.21 (d, J = 8.0 Hz, 2H), 7.09 (d, J = 7.2 Hz, 2H), 5.13 (s, 1H), 4.78 (d, J = 13.6 Hz, 1H), 4.45 (d, J = 10.8 Hz, 1H), 2.89–2.79 (m,1H), 1.20 (d, J = 6.8 Hz, 6H). 13C NMR (151 MHz, CDCl3) δ 149.03, 132.73, 132.53, 132.04, 131.31, 131.14, 130.76, 130.47, 129.81, 129.35, 129.27, 128.61, 128.28, 128.21, 126.84, 75.89, 75.71, 45.77, 45.33, 33.68, 23.82. 31P NMR (243 MHz, CDCl3) δ 30.26.

3.5.5. 3e ((1-(4-fluorophenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, DMSO-d6) δ 8.20–8.08 (m, 2H), 7.74 (ddd, J = 11.6, 6.8, 1.6 Hz, 2H), 7.69–7.63 (m, 3H), 7.51–7.33 (m, 5H), 7.06 (t, J = 8.8 Hz, 2H), 5.12–5.07 (m, 2H), 4.85 (s, 1H). 13C NMR (151 MHz, DMSO) δ 139.74, 139.53, 133.09, 132.49, 131.88, 131.31, 131.25, 131.09, 131.04, 130.74, 130.39, 129.73, 129.65, 129.31, 128.93, 128.86, 115.70, 115.55, 76.24, 76.23, 43.56, 43.14. 31P NMR (243 MHz, DMSO) δ 29.97. 19F NMR (565 MHz, DMSO) δ -114.23.

3.5.6. 3f ((1-(4-chlorophenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, Chloroform-d) δ 7.99 (dd, J = 11.2, 7.6 Hz, 2H), 7.70–7.59 (m, 3H), 7.53–7.41 (m, 3H), 7.34 (dt, J = 10.8, 5.2 Hz, 2H), 7.27 (dd, J = 8.4, 1.6 Hz, 2H), 7.21 (d, J = 8.4 Hz, 2H), 5.06 (d, J = 3.2 Hz, 1H), 4.78–4.70 (m, 1H), 4.47–4.38 (m, 1H). 13C NMR (151 MHz, CDCl3) δ 134.43, 132.92, 132.34, 131.15, 131.09, 130.95, 130.89, 130.74, 130.32, 130.13, 129.48, 129.40, 129.05, 128.59, 128.51, 75.69, 75.65, 45.46, 45.04. 31P NMR (243 MHz, CDCl3) δ 29.53.

3.5.7. 3g ((1-(4-bromophenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, DMSO-d6) δ 8.13 (td, J = 8.0, 3.6 Hz, 2H), 7.76 (dd, J = 11.6, 7.6 Hz, 2H), 7.69–7.62 (m, 3H), 7.48–7.35 (m, 7H), 5.13–5.05 (m, 2H), 4.90–4.80 (m, 1H). 13C NMR (151 MHz, DMSO) δ 133.13, 132.71, 132.60, 132.19, 131.99, 131.67, 131.29, 131.10, 131.03, 130.65, 130.27, 129.75, 129.67, 128.94, 121.60, 76.00, 43.75, 43.33. 31P NMR (243 MHz, DMSO) δ 29.69.

3.5.8. 3h ((1-(2-chlorophenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, Chloroform-d) δ 8.10 (ddd, J = 11.2, 7.6, 1.6 Hz, 2H), 7.89 (d, J = 7.6 Hz, 1H), 7.74–7.64 (m, 3H), 7.49–7.39 (m, 3H), 7.34 (dt, J = 8.4, 4.0 Hz, 1H), 7.26 (dd, J = 7.6, 3.2 Hz, 2H), 7.19 (d, J = 4.0 Hz, 2H), 5.25 (dd, J = 9.2, 3.2 Hz, 1H), 5.20–5.11 (m, 1H), 4.79 (ddd, J = 13.6, 6.0, 3.2 Hz, 1H). 13C NMR (151 MHz, CDCl3) δ 134.95, 132.97, 132.26, 131.94, 131.30, 131.24, 130.95, 130.89, 130.19, 129.91, 129.71, 129.56, 129.47, 129.41, 128.42, 128.20, 128.12, 127.52, 75.39, 75.38, 41.14, 40.72. 31P NMR (243 MHz, CDCl3) δ 30.73.

3.5.9. 3i ((1-(3-chlorophenyl)-2-nitroethyl) diphenylphosphine oxide)

1H NMR (400 MHz, Chloroform-d) δ 8.01 (dd, J = 11.2, 7.6 Hz, 2H), 7.67 (t, J = 9.2 Hz, 3H), 7.53–7.43 (m, 3H), 7.36 (dt, J = 8.8, 4.4 Hz, 2H), 7.28–7.16 (m, 4H), 5.15–5.03 (m, 1H), 4.80–4.73 (m, 1H), 4.46–4.36 (m, 1H). 13C NMR (151 MHz, CDCl3) δ 134.61, 133.83, 132.98, 132.41, 131.19, 131.13, 130.99, 130.93, 130.01, 129.65, 129.49, 129.42, 128.57, 128.49, 127.47, 73.47, 73.45, 45.76, 45.34. 31P NMR (243 MHz, CDCl3) δ 29.82.

3.5.10. 3j (1-(3,4-dimethoxyphenyl)-2-nitroethyl) diphenylphosphine oxide

1H NMR (400 MHz, Chloroform-d) δ 7.99 (dd, J = 11.2, 7.6 Hz, 2H), 7.70–7.59 (m, 3H), 7.53–7.41 (m, 3H), 7.34 (dt, J = 10.8, 5.2 Hz, 2H), 7.27 (dd, J = 8.4, 1.6 Hz, 2H), 7.21 (d, J = 8.4 Hz, 2H), 5.06 (d, J = 3.2 Hz, 1H), 4.78–4.70 (m, 1H), 4.47–4.38 (m, 1H), 3.88 (s, 3H), 3.77 (s, 3H). 13C NMR (151 MHz, CDCl3) δ 134.43, 132.92, 132.34, 131.15, 131.09, 130.95, 130.89, 130.74, 130.32, 130.13, 129.47, 129.40, 129.04, 128.59, 128.51, 122.45, 122.05, 122.00, 75.67, 55.45, 55.45, 45.46, 45.04. 31P NMR (243 MHz, CDCl3) δ 29.53.

3.5.11. 5a (2-((diphenylphosphoryl)(phenyl)methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.05–7.97 (m, 2H), 7.72–7.60 (m, 3H), 7.56–7.49 (m, 2H), 7.45–7.39 (m, 3H), 7.31 (td, J = 5.2, 4.8, 3.2 Hz, 5H), 4.78 (t, J = 7.6 Hz, 1H), 4.09 (s, 1H). 13C NMR (151 MHz, CDCl3) δ 134.43, 132.92, 132.34, 131.15, 131.09, 130.95, 130.89, 130.74, 130.32, 130.13, 129.47, 129.40, 129.04, 128.59, 128.51, 111.14, 111.09, 111.05, 75.67, 45.46, 45.04, 24.31. 31P NMR (243 MHz, CDCl3) δ 29.53.

3.5.12. 5b (2-((2-chlorophenyl) (diphenylphosphoryl)methyl) malononitrile)

1H NMR (400 MHz, DMSO-d6) δ 8.20–8.12 (m, 2H), 7.99 (d, J = 8.0 Hz, 1H), 7.76–7.64 (m, 3H), 7.50 (dd, J = 11.6, 7.6 Hz, 2H), 7.42 (t, J = 7.2 Hz, 2H), 7.36–7.26 (m, 4H), 5.66 (dd, J = 8.4, 6.8 Hz, 1H), 5.22–5.15 (m, 1H). 13C NMR (151 MHz, DMSO) δ 134.63, 134.19, 133.99, 133.45, 132.74, 132.28, 132.20, 131.84, 131.78, 130.97, 130.76, 130.28, 130.14, 129.68, 129.60, 129.24, 129.16, 128.81, 128.73, 113.20, 113.13, 112.01, 111.91, 42.11, 41.83, 25.05. 31P NMR (243 MHz, DMSO) δ 29.01.

3.5.13. 5c (2-((3-chlorophenyl) (diphenylphosphoryl)methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.02 (dd, J = 11.6, 7.6 Hz, 2H), 7.75–7.63 (m, 3H), 7.59–7.44 (m, 3H), 7.36 (d, J = 7.2 Hz, 5H), 7.27 (d, J = 8.0 Hz, 1H), 4.74 (s, 1H), 4.03 (t, J = 7.6 Hz, 1H). 13C NMR (151 MHz, CDCl3) δ 135.15, 133.43, 131.32, 131.30, 131.20, 131.18, 130.53, 129.87, 129.65, 129.48, 129.41, 128.65, 128.57, 127.72, 110.94, 110.93, 47.07, 46.67, 24.62. 31P NMR (243 MHz, CDCl3) δ 28.36.

3.5.14. 5d (2-((diphenylphosphoryl)(3-nitrophenyl) methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.19 (d, J = 8.8 Hz, 2H), 8.09–8.02 (m, 2H), 7.96 (d, J = 7.6 Hz, 1H), 7.71 (ddd, J = 18.8, 7.6, 2.4 Hz, 3H), 7.57 (ddd, J = 11.2, 8.0, 2.0 Hz, 3H), 7.44 (d, J = 7.2 Hz, 1H), 7.35 (td, J = 7.6, 3.2 Hz, 2H), 4.75 (s, 1H), 4.20 (t, J = 7.2 Hz, 1H). 13C NMR (151 MHz, CDCl3) δ 135.18, 133.71, 133.09, 132.91, 131.33, 131.27, 130.95, 130.89, 130.47, 129.67, 129.59, 128.91, 128.83, 125.11, 124.25, 110.68, 46.98, 46.57, 24.47. 31P NMR (243 MHz, CDCl3) δ 28.24.

3.5.15. 5e (2-((diphenylphosphoryl)(m-tolyl) methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.03–7.97 (m, 2H), 7.66 (dtd, J = 14.8, 7.2, 2.4 Hz, 3H), 7.55–7.48 (m, 2H), 7.43 (dd, J = 7.6, 1.6 Hz, 1H), 7.33 (dd, J = 7.6, 3.6 Hz, 2H), 7.22–7.11 (m, 4H), 4.83 (s, 1H), 4.03 (t, J = 8.0 Hz, 1H), 2.28 (s, 3H). 13C NMR (151 MHz, CDCl3) δ 133.17, 132.69, 132.37, 131.36, 130.58, 130.30, 130.13, 129.30, 129.22, 129.12, 128.88, 128.35, 128.27, 126.77, 111.31, 111.30, 47.42, 47.00, 24.70, 21.30. 31P NMR (243 MHz, CDCl3) δ 28.64.

3.5.16. 5f (2-((diphenylphosphoryl)(3-methoxyphenyl) methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.04–7.95 (m, 2H), 7.71–7.53 (m, 5H), 7.45 (d, J = 1.6 Hz, 1H), 7.33 (d, J = 3.2 Hz, 2H), 7.22 (d, J = 8.0 Hz, 1H), 7.01–6.92 (m, 2H), 6.86 (dd, J = 8.4, 2.4 Hz, 1H), 4.77 (s, 1H), 4.05 (t, J = 8.0 Hz, 1H), 3.73 (s, 3H). 13C NMR (151 MHz, CDCl3) δ 133.20, 132.46, 132.03, 131.36, 131.36, 130.31, 129.65, 129.35, 129.27, 128.85, 128.45, 128.46, 128.38, 122.08, 115.48, 114.83, 111.28, 55.29, 47.43, 47.01, 24.78. 31P NMR (243 MHz, CDCl3) δ 30.84.

3.5.17. 5g 2-((diphenylphosphoryl)(4-fluorophenyl) methyl) malononitrile

1H NMR (400 MHz, DMSO-d6) δ 8.17–8.09 (m, 2H), 7.69–7.62 (m, 5H), 7.53 (s, 4H), 7.44–7.32 (m, 3H), 5.46 (t, J = 7.2 Hz, 1H), 5.17 (t, J = 7.2 Hz, 1H). 13C NMR (151 MHz, DMSO) δ 133.11, 132.60, 132.46, 132.39, 132.32, 131.88, 131.62, 131.56, 131.37, 131.30, 130.82, 130.76, 130.16, 130.10, 129.49, 129.42, 128.87, 128.79, 116.00, 115.85, 115.09, 114.95, 113.17, 112.93, 79.57, 79.35, 79.14, 42.23, 41.80, 25.47. 31P NMR (243 MHz, DMSO) δ 28.58. 19F NMR (565 MHz, DMSO) δ -112.90.

3.5.18. 5h 2-((4-bromophenyl) (diphenylphosphoryl)methyl) malononitrile

1H NMR (400 MHz, Chloroform-d) δ 8.03–7.97 (m, 2H), 7.83–7.50 (m, 7H), 7.44 (dtd, J = 8.4, 3.9, 2.0 Hz, 3H), 7.35 (td, J = 7.6, 3.2 Hz, 1H), 7.02 (t, J = 8.6 Hz, 1H), 4.69 (t, J = 7.2 Hz, 1H), 4.05 (t, J = 7.6 Hz, 1H). 13C NMR (151 MHz, CDCl3) δ 137.03, 136.79, 133.15, 132.51, 131.97, 131.28, 130.69, 130.43, 130.25, 129.81, 129.27, 128.80, 128.45, 111.89, 111.37, 78.00, 77.95, 42.23, 41.80, 25.47. 31P NMR (243 MHz, CDCl3) δ 29.69.

3.5.19. 5i (2-((3,4-dimethoxyphenyl) (diphenylphosphoryl)methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.02–7.92 (m, 2H), 7.72–7.53 (m, 5H), 7.47 (t, J = 7.6 Hz, 1H), 7.36 (td, J = 7.6, 2.8 Hz, 2H), 6.96 (d, J = 5.6 Hz, 2H), 6.82 (d, J = 8.4 Hz, 1H), 4.73 (s, 1H), 4.04–3.96 (m, 1H), 3.88 (s, 3H), 3.77 (s, 3H). 13C NMR (151 MHz, CDCl3) δ 139.74, 139.53, 133.21, 132.49, 131.88, 131.31, 131.25, 131.09, 131.04, 130.74, 130.39, 129.73, 129.65, 129.31, 128.94, 128.86, 78.00, 77.95, 57.37, 56.85, 47.42, 47.01, 24.70. 31P NMR (243 MHz, CDCl3) δ 28.47.

3.5.20. 5j (2-((diphenylphosphoryl)(naphthalen-2-yl) methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 8.03 (dd, J = 11.6, 7.6 Hz, 2H), 7.91 (s, 1H), 7.81 (t, J = 9.2 Hz, 4H), 7.73–7.61 (m, 3H), 7.58–7.46 (m, 5H), 7.25 (td, J = 7.6, 3.2 Hz, 2H), 4.85 (t, J = 7.6 Hz, 1H), 4.23 (t, J = 8.0 Hz, 1H). 13C NMR (151 MHz, CDCl3) δ 133.23, 133.11, 132.45, 132.02, 131.35, 131.29, 130.38, 129.72, 129.49, 129.38, 129.31, 129.26, 128.79, 128.46, 128.38, 128.18, 127.72, 127.07, 126.80, 126.48, 111.36, 111.30, 111.21, 47.54, 47.12, 24.97. 31P NMR (243 MHz, CDCl3) δ 28.51.

3.5.21. 5k (2-((diphenylphosphoryl)(thiophen-2-yl) methyl) malononitrile)

1H NMR (400 MHz, Chloroform-d) δ 7.96 (q, J = 7.6, 6.4 Hz, 2H), 7.86–7.74 (m, 2H), 7.70 (d, J = 7.2 Hz, 1H), 7.64 (d, J = 7.6 Hz, 4H), 7.52 (s, 1H), 7.41 (d, J = 7.2 Hz, 2H), 7.25 (d, J = 3.2 Hz, 1H), 4.75 (s, 1H), 4.40 (s, 1H). 13C NMR (151 MHz, CDCl3) δ 133.37, 132.78, 132.35, 131.89, 131.64, 131.58, 131.41, 131.34, 130.80, 130.68, 129.81, 129.42, 129.34, 129.16, 128.87, 128.55, 128.47, 128.36, 128.16, 127.85, 127.45, 126.82, 111.25, 111.19, 110.96, 110.92, 42.97, 42.54, 25.62. 31P NMR (243 MHz, CDCl3) δ 28.36.

4. Conclusions

In conclusion, we developed a lipase-catalyzed phospha-Michael addition reaction of β-nitrostyrenes or benzylidenemalononitrile with diphenylphosphine oxide as a phosphine source under mild reaction conditions. This enzymatic method utilizes available substrates, environment-friendly solvents, and simple and practical operation (requiring only filtration and washing without column chromatography). Novozym 435 has been demonstrated to be an efficient, atom-economic, and reusable catalyst for the synthesis of various organophosphorus compounds. Furthermore, immobilization is an efficient tool for improving enzyme features in the biotechnology toolbox. To improve the feasibility and efficiency of this synthetic method, other Cal-B immobilization methods are being studied and will be reported in due course [58,59,60].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27227798/s1, Figure S1: The time course of different dosage lipase-catalyzed phospha-Michael addition reaction; Table S1: Reusability of Novozym 435 on the phospha-Michael addition reaction.

Author Contributions

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

Funding

This research was funded by the Science and Technology Development Program of Jilin Province (No. 20200301029RQ, No. 20200201396JC) and the Key Research Development Program of Jilin Province (No. 20200402067NC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Samples are not available from the corresponding authors.

Acknowledgments

We gratefully acknowledge the Science and Technology Development Program of Jilin Province (No. 20200301029RQ, No. 20200201396JC) and the Key Research Development Program of Jilin Province (No. 20200402067NC).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Yu, H.; Yang, H.; Shi, E.; Tang, W. Development and Clinical Application of Phosphorus-Containing Drugs. Med. Drug Discov. 2020, 8, 100063. [Google Scholar] [CrossRef] [PubMed]
  2. Shevchuk, M.; Wang, Q.; Pajkert, R.; Xu, J.; Mei, H.; Röschenthaler, G.-V.; Han, J. Recent Advances in Synthesis of Difluoromethylene Phosphonates for Biological Applications. Adv. Synth. Catal. 2021, 363, 2912–2968. [Google Scholar] [CrossRef]
  3. Fields, S.C. Synthesis of natural products containing a C-P bond. Tetrahedron 1999, 55, 12237–12273. [Google Scholar] [CrossRef]
  4. Walsh, C.T.; Tu, B.P.; Tang, Y. Eight Kinetically Stable but Thermodynamically Activated Molecules that Power Cell Metabolism. Chem. Rev. 2018, 118, 1460–1494. [Google Scholar] [CrossRef]
  5. Fernández-García, C.; Coggins, A.J.; Powner, M.W. A Chemist’s Perspective on the Role of Phosphorus at the Origins of Life. Life 2017, 7, 31. [Google Scholar] [CrossRef] [Green Version]
  6. Miller, R.C.; Bradley, J.S.; Hamilton, L.A. Disubstituted Phosphine Oxides. III. Addition to α,β-Unsaturated Nitriles and Carbonyl Compounds1. J. Am. Chem. Soc. 1956, 78, 5299–5303. [Google Scholar] [CrossRef]
  7. Bodalski, R.; Pietrusiewicz, K. A new route to the phospholane ring-system. Tetrahedron Lett. 1972, 13, 4209–4212. [Google Scholar] [CrossRef]
  8. Enders, D.; Mirjafary, Z.; Saeidian, H. Efficient diastereo- and enantioselective synthesis of α,β-disubstituted γ-phosphono sulfonates. Tetrahedron Asymmetry 2009, 20, 2429–2431. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Wang, W. Recent advances in organocatalytic asymmetric Michael reactions. Catal. Sci. Technol. 2012, 2, 42–53. [Google Scholar] [CrossRef]
  10. Strappaveccia, G.; Bianchi, L.; Ziarelli, S.; Santoro, S.; Lanari, D.; Pizzo, F.; Vaccaro, L. PS-BEMP as a basic catalyst for the phospha-Michael addition to electron-poor alkenes. Org. Biomol. Chem. 2016, 14, 3521–3525. [Google Scholar] [CrossRef]
  11. Lenz, J.; Pospiech, D.; Komber, H.; Paven, M.; Albach, R.; Mentizi, S.; Langstein, G.; Voit, B. Synthesis of the H-phosphonate dibenzo[d,f][1,3,2]dioxaphosphepine 6-oxide and the phospha-Michael addition to unsaturated compounds. Tetrahedron 2019, 75, 1306–1310. [Google Scholar] [CrossRef]
  12. Tessema, E.; Elakkat, V.; Chiu, C.-F.; Zheng, J.-H.; Chan, K.L.; Shen, C.-R.; Zhang, P.; Lu, N. Recoverable Phospha-Michael Additions Catalyzed by a 4-N,N-Dimethylaminopyridinium Saccharinate Salt or a Fluorous Long-Chained Pyridine: Two Types of Reusable Base Catalysts. Molecules 2021, 26, 1159. [Google Scholar] [CrossRef] [PubMed]
  13. Trepohl, V.T.; Mori, S.; Itami, K.; Oestreich, M. Palladium(II)-Catalyzed Conjugate Phosphination of Electron-Deficient Acceptors. Org. Lett. 2009, 11, 1091–1094. [Google Scholar] [CrossRef] [PubMed]
  14. Trepohl, V.T.; Fröhlich, R.; Oestreich, M. Conjugate phosphination of cyclic and acyclic acceptors using Rh(I)–phosphine or Rh(I)–carbene complexes. Probing the mechanism with chirality at the silicon atom or the phosphorus atom of the Si–P reagent. Tetrahedron 2009, 65, 6510–6518. [Google Scholar] [CrossRef]
  15. Yan, J.; Wang, Y.-B.; Hou, S.; Shi, L.; Zhu, X.; Hao, X.-Q.; Song, M.-P. NCC Pincer Ni (II) Complexes Catalyzed Hydrophosphination of Nitroalkenes with Diphenylphosphine. Appl. Organomet. Chem. 2020, 34, e5954. [Google Scholar] [CrossRef]
  16. Hao, X.-Q.; Huang, J.-J.; Wang, T.; Lv, J.; Gong, J.-F.; Song, M.-P. PCN Pincer Palladium(II) Complex Catalyzed Enantioselective Hydrophosphination of Enones: Synthesis of Pyridine-Functionalized Chiral Phosphine Oxides as NCsp3O Pincer Preligands. J. Org. Chem 2014, 79, 9512–9530. [Google Scholar] [CrossRef]
  17. Miao, H.; Wang, S.; Zhu, X.; Zhou, S.; Wei, Y.; Yuan, Q.; Mu, X. Synthesis, characterization and catalytic activity of rare-earth metal amides incorporating cyclohexyl bridged bis(β-diketiminato) ligands. RSC Adv. 2017, 7, 42792–42799. [Google Scholar] [CrossRef] [Green Version]
  18. Ding, B.; Zhang, Z.; Xu, Y.; Liu, Y.; Sugiya, M.; Imamoto, T.; Zhang, W. P-Stereogenic PCP Pincer–Pd Complexes: Synthesis and Application in Asymmetric Addition of Diarylphosphines to Nitroalkenes. Org. Lett. 2013, 15, 5476–5479. [Google Scholar] [CrossRef]
  19. Feng, J.-J.; Huang, M.; Lin, Z.-Q.; Duan, W.-L. Palladium-Catalyzed Asymmetric 1,4-Addition of Diarylphosphines to Nitroalkenes for the Synthesis of Chiral P,N-Compounds. Adv. Synth. Catal. 2012, 354, 3122–3126. [Google Scholar] [CrossRef]
  20. Wani, A.A.; Chourasiya, S.S.; Kathuria, D.; Bharatam, P.V. 1,1-Diaminoazines as organocatalysts in phospha-Michael addition reactions. Chem. Commun. 2021, 57, 11717–11720. [Google Scholar] [CrossRef]
  21. Fu, X.; Jiang, Z.; Tan, C.-H. Bicyclic guanidine-catalyzed enantioselective phospha-Michael reaction: Synthesis of chiral β-aminophosphine oxides and β-aminophosphines. Chem. Commun. 2007, 47, 5058–5060. [Google Scholar] [CrossRef] [PubMed]
  22. Jiang, Z.; Zhang, Y.; Ye, W.; Tan, C.-H. P–C Bond formation via direct and three-component conjugate addition catalyzed by 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD). Tetrahedron Lett. 2007, 48, 51–54. [Google Scholar] [CrossRef]
  23. Xu, Y.; Li, F.; Zhao, N.; Su, J.; Wang, C.; Wang, C.; Li, Z.; Wang, L. Environment-friendly and efficient synthesis of 2-aminobenzo-xazoles and 2-aminobenzothiazoles catalyzed by Vitreoscilla hemoglobin incorporating a cobalt porphyrin cofactor. Green Chem. 2021, 23, 8047–8052. [Google Scholar] [CrossRef]
  24. 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]
  25. Hanefeld, U.; Hollmann, F.; Paul, C.E. Biocatalysis making waves in organic chemistry. Chem. Soc. Rev. 2022, 51, 594–627. [Google Scholar] [CrossRef] [PubMed]
  26. Peng, Y.; Gao, C.; Zhang, Z.; Wu, S.; Zhao, J.; Li, A. A Chemoenzymatic Strategy for the Synthesis of Steroid Drugs Enabled by P450 Monooxygenase-Mediated Steroidal Core Modification. ACS Catal. 2022, 12, 2907–2914. [Google Scholar] [CrossRef]
  27. Sheldon, R.A.; Brady, D. Broadening the Scope of Biocatalysis in Sustainable Organic Synthesis. ChemSusChem 2019, 12, 2859–2881. [Google Scholar] [CrossRef] [PubMed]
  28. Arnold, F.H. Innovation by Evolution: Bringing New Chemistry to Life (Nobel Lecture). Angew. Chem. Int. Ed. 2019, 58, 14420–14426. [Google Scholar] [CrossRef] [Green Version]
  29. Bornscheuer, U.T.; Kazlauskas, R.J. Catalytic promiscuity in biocatalysis: Using old enzymes to form new bonds and follow new pathways. Angew. Chem. Int. Ed. 2004, 43, 6032–6040. [Google Scholar] [CrossRef]
  30. Woodley, J.M. Integrating protein engineering into biocatalytic process scale-up. Trends Chem. 2022, 4, 371–373. [Google Scholar] [CrossRef]
  31. Arana-Peña, S.; Rios, N.S.; Carballares, D.; Goncalves, L.R.; Fernandez-Lafuente, R. Immobilization of lipases via interfacial activation on hydrophobic supports: Production of biocatalysts libraries by altering the immobilization conditions. Catal. Today 2021, 362, 130–140. [Google Scholar] [CrossRef]
  32. Sheldon, R.A.; Brady, D.; Bode, M.L. The Hitchhiker's guide to biocatalysis: Recent advances in the use of enzymes in organic synthesis. Chem. Sci. 2020, 11, 2587–2605. [Google Scholar] [CrossRef] [Green Version]
  33. Dwivedee, B.P.; Soni, S.; Sharma, M.; Bhaumik, J.; Laha, J.K.; Banerjee, U.C. Promiscuity of lipase-catalyzed reactions for organic synthesis: A recent update. ChemistrySelect 2018, 3, 2441–2466. [Google Scholar] [CrossRef]
  34. Li, F.; Xu, Y.; Wang, C.; Wang, C.; Zhao, R.; Wang, L. Efficient synthesis of cyano-containing multi-substituted indoles catalyzed by lipase. Bioorg. Chem. 2021, 107, 104583. [Google Scholar] [CrossRef]
  35. Zhao, Z.; Zhang, L.; Li, F.; Tang, X.; Ma, Y.; Wang, C.; Wang, Z.; Zhao, R.; Wang, L. A Novel Oxidation of Salicyl Alcohols Catalyzed by Lipase. Catalysts 2017, 7, 354. [Google Scholar] [CrossRef] [Green Version]
  36. Li, C.; Feng, X.-W.; Wang, N.; Zhou, Y.-J.; Yu, X.-Q. Biocatalytic promiscuity: The first lipase-catalysed asymmetric aldol reaction. Green Chem. 2008, 10, 616–618. [Google Scholar] [CrossRef]
  37. Bavandi, H.; Habibi, Z.; Yousefi, M. Porcine pancreas lipase as a green catalyst for synthesis of bis-4-hydroxy coumarins. Bioorg. Chem. 2020, 103, 104139. [Google Scholar] [CrossRef] [PubMed]
  38. Xu, F.; Wu, Q.; Chen, X.; Lin, X.; Wu, Q. A Single Lipase-Catalysed One-Pot Protocol Combining Aminolysis Resolution and Aza-Michael Addition: An Easy and Efficient Way to Synthesise β-Amino Acid Esters. Eur. J. Org. Chem. 2015, 2015, 5393–5401. [Google Scholar] [CrossRef]
  39. González-Martínez, D.; Gotor, V.; Gotor-Fernández, V. Application of Deep Eutectic Solvents in Promiscuous Lipase-Catalysed Aldol Reactions. Eur. J. Org. Chem. 2016, 2016, 1513–1519. [Google Scholar] [CrossRef]
  40. Mathebula, N.P.; Sheldon, R.A.; Bode, M.L. Lipase-Catalysed Enzymatic Kinetic Resolution of Aromatic Morita-Baylis-Hillman Derivatives by Hydrolysis and Transesterification. ChemBioChem 2022, 23, e202200435. [Google Scholar] [CrossRef]
  41. Samsonowicz-Górski, J.; Koszelewski, D.; Kowalczyk, P.; Śmigielski, P.; Hrunyk, A.; Kramkowski, K.; Wypych, A.; Szymczak, M.; Lizut, R.; Ostaszewski, R. Promiscuous Lipase-Catalyzed Knoevenagel–Phospha–Michael Reaction for the Synthesis of Antimicrobial β-Phosphono Malonates. Int. J. Mol. Sci. 2022, 23, 8819. [Google Scholar] [CrossRef]
  42. Priego, J.; Ortíz-Nava, C.; Carrillo-Morales, M.; López-Munguía, A.; Escalante, J.; Castillo, E. Solvent engineering: An effective tool to direct chemoselectivity in a lipase-catalyzed Michael addition. Tetrahedron 2009, 65, 536–539. [Google Scholar] [CrossRef]
  43. Rivera-Ramírez, J.D.; Escalante, J.; López-Munguía, A.; Marty, A.; Castillo, E. Thermodynamically controlled chemoselectivity in lipase-catalyzed aza-Michael additions. J. Mol. Catal. 2015, 112, 76–82. [Google Scholar] [CrossRef]
  44. Yang, F.; Wang, Z.; Zhang, X.; Jiang, L.; Li, Y.; Wang, L. A Green Chemoenzymatic Process for the Synthesis of Azoxybenzenes. ChemCatChem 2015, 7, 3450–3453. [Google Scholar] [CrossRef]
  45. Zhang, J.; Wang, C.; Wang, C.; Shang, W.; Xiao, B.; Duan, S.; Li, F.; Wang, L.; Chen, P. Lipase-catalyzed aza-Michael addition of amines to acrylates in supercritical carbon dioxide. J. Chem. Technol. Biotechnol. 2019, 94, 3981–3986. [Google Scholar] [CrossRef]
  46. Klibanov, A.M. Improving enzymes by using them in organic solvents. Nature 2001, 409, 241–246. [Google Scholar] [CrossRef]
  47. 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] [Green Version]
  48. Spahn, C.; Minteer, D.S. Enzyme Immobilization in Biotechnology. Recent Pat. Eng. 2008, 2, 195–200. [Google Scholar] [CrossRef]
  49. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzyme Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  50. Sampaio, C.S.; Angelotti, J.A.F.; Fernandez-Lafuente, R.; Hirata, D.B. Lipase immobilization via cross-linked enzyme aggregates: Problems and prospects—A review. Int. J. Biol. Macromol. 2022, 215, 434–449. [Google Scholar] [CrossRef]
  51. Morellon-Sterling, R.; Tavano, O.; Bolivar, J.M.; Berenguer-Murcia, Á.; Vela-Gutiérrez, G.; Sabir, J.S.M.; Tacias-Pascacio, V.G.; Fernandez-Lafuente, R. A review on the immobilization of pepsin: A Lys-poor enzyme that is unstable at alkaline pH values. Int. J. Biol. Macromol. 2022, 210, 682–702. [Google Scholar] [CrossRef]
  52. de Albuquerque, T.L.; de Sousa, M.; Gomes e Silva, N.C.; Girão Neto, C.A.C.; Gonçalves, L.R.B.; Fernandez-Lafuente, R.; Rocha, M.V.P. β-Galactosidase from Kluyveromyces lactis: Characterization, production, immobilization and applications—A review. Int. J. Biol. Macromol. 2021, 191, 881–898. [Google Scholar] [CrossRef]
  53. Altinkaynak, C.; Tavlasoglu, S.; ÿzdemir, N.; Ocsoy, I. A new generation approach in enzyme immobilization: Organic-inorganic hybrid nanoflowers with enhanced catalytic activity and stability. Enzyme Microb. Technol. 2016, 93–94, 105–112. [Google Scholar] [CrossRef]
  54. Palomo, J.M.; Muñoz, G.; Fernández-Lorente, G.; Mateo, C.; Fernández-Lafuente, R.; Guisán, J.M. Interfacial adsorption of lipases on very hydrophobic support (octadecyl–Sepabeads): Immobilization, hyperactivation and stabilization of the open form of lipases. J. Mol. Catal. B Enzym. 2002, 19–20, 279–286. [Google Scholar] [CrossRef]
  55. 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] [PubMed]
  56. Verdasco-Martín, C.M.; Garcia-Verdugo, E.; Porcar, R.; Fernandez-Lafuente, R.; Otero, C. Selective synthesis of partial glycerides of conjugated linoleic acids via modulation of the catalytic properties of lipases by immobilization on different supports. Food Chem. 2018, 245, 39–46. [Google Scholar] [CrossRef] [PubMed]
  57. Esteve, F.; Escrig, A.; Porcar, R.; Luis, S.V.; Altava, B.; García-Verdugo, E. Immobilized Supramolecular Systems as Efficient Synzymes for CO2 Activation and Conversion. Adv. Sustain. Syst. 2022, 6, 2100408. [Google Scholar] [CrossRef]
  58. Cipolatti, E.P.; Rios, N.S.; Sousa, J.S.; Robert, J.d.M.; da Silva, A.A.T.; Pinto, M.C.C.; Simas, A.B.C.; Vilarrasa-García, E.; Fernandez-Lafuente, R.; Gonçalves, L.R.B.; et al. Synthesis of lipase/silica biocatalysts through the immobilization of CALB on porous SBA-15 and their application on the resolution of pharmaceutical derivatives and on nutraceutical enrichment of natural oil. Mol. Catal. 2021, 505, 111529. [Google Scholar] [CrossRef]
  59. Arana-Peña, S.; Carballares, D.; Morellon-Sterling, R.; Rocha-Martin, J.; Fernandez-Lafuente, R. The combination of covalent and ionic exchange immobilizations enables the coimmobilization on vinyl sulfone activated supports and the reuse of the most stable immobilized enzyme. Int. J. Biol. Macromol. 2022, 199, 51–60. [Google Scholar] [CrossRef]
  60. Arana-Peña, S.; Rios, N.S.; Mendez-Sanchez, C.; Lokha, Y.; Gonçalves, L.R.B.; Fernández-Lafuente, R. Use of polyethylenimine to produce immobilized lipase multilayers biocatalysts with very high volumetric activity using octyl-agarose beads: Avoiding enzyme release during multilayer production. Enzyme Microb. Technol. 2020, 137, 109535. [Google Scholar] [CrossRef]
Scheme 1. Previous work’s synthesis for phospha-Michael addition reactions and synthesis of C-P bond compounds catalyzed by lipase.
Scheme 1. Previous work’s synthesis for phospha-Michael addition reactions and synthesis of C-P bond compounds catalyzed by lipase.
Molecules 27 07798 sch001
Figure 1. Reusability of Novozym 435 in enzymatic synthesis. Reaction conditions: 1a (0.5 mmol), 2 (0.5 mmol), Novozym 435 (300 U) rt, and EtOH (1 mL).
Figure 1. Reusability of Novozym 435 in enzymatic synthesis. Reaction conditions: 1a (0.5 mmol), 2 (0.5 mmol), Novozym 435 (300 U) rt, and EtOH (1 mL).
Molecules 27 07798 g001
Scheme 2. Plausible mechanism for the phospha-Michael addition reactions.
Scheme 2. Plausible mechanism for the phospha-Michael addition reactions.
Molecules 27 07798 sch002
Table 1. Lipase-catalyzed phospha-Michael addition reaction along with the reaction optimization details.
Table 1. Lipase-catalyzed phospha-Michael addition reaction along with the reaction optimization details.
Molecules 27 07798 i001
EntrySolventCatalyst 2TimeYield (%) 1
1EtOHNovozym 4352 h94
2EtOHPPL4 h89
3EtOHCal-B4 h78
4EtOHCSL4 h64
5EtOHTEA (20 mol%)12 h30
6EtOHNo catalyst12 hN. D. 3
7EtOHNovozym 435 412 hN. D.
8EtOHNovozym 435 512 hN. D.
9MeOHNovozym 4352 h78
10MeCNNovozym 4352 h72
11CH2Cl2Novozym 4352 h42
12DioxaneNovozym 4352 h12
13THFNovozym 4352 h28
14TolueneNovozym 4352 h32
15hexaneNovozym 4352 h25
16WaterNovozym 4352 h40
17EtOHNovozym 435 (150 U)2 h81
18EtOHNovozym 435 (600 U)2 h95
Reaction conditions: 1a (R = H, 0.5 mmol), 2 (0.5 mmol), lipase (300 U), rt, and solvent (1 mL). 1 Isolated yield. 2 PPL (Porcine pancreatic lipase); Cal-B (C. antarctica lipase B); CSL (Candida sp. lipase); PSL (Lipase from Pseudomonas sp.). 3 Not detected. 4 Novozym 435 was inactivated by phenylmethanesulfonyl fluoride (PMSF). 5 Thermally deactivated Novozym 435 boiled for 12 h in water was used as the catalyst.
Table 2. Substrate scope of phospha-Michael addition reactions (1).
Table 2. Substrate scope of phospha-Michael addition reactions (1).
Molecules 27 07798 i002
Molecules 27 07798 i003
Reaction conditions: 1 (0.5 mmol), 2 (0.5 mmol), Novozym435 (300 U), rt, and EtOH (1 mL).
Table 3. Substrate scope of phospha-Michael addition reactions (4).
Table 3. Substrate scope of phospha-Michael addition reactions (4).
Molecules 27 07798 i004
Molecules 27 07798 i005
Reaction conditions: 4 (0.5 mmol), 2 (0.5 mmol), Novozym 435 (300 U) rt, and EtOH (1 mL).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/molecules27227798

AMA Style

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(22):7798. https://doi.org/10.3390/molecules27227798

Chicago/Turabian Style

Xu, Yuelin, Fengxi Li, Jinglin Ma, Jiapeng Li, Hanqing Xie, Chunyu Wang, Peng Chen, and Lei Wang. 2022. "Lipase-Catalyzed Phospha-Michael Addition Reactions under Mild Conditions" Molecules 27, no. 22: 7798. https://doi.org/10.3390/molecules27227798

Article Metrics

Back to TopTop