Next Article in Journal
Characterization and Performance of Peanut Shells in Caffeine and Triclosan Removal in Batch and Fixed-Bed Column Tests
Previous Article in Journal
Rapid Determination of Tetracyclines in Drinking and Environmental Waters Using Fully Automatic Solid-Phase Extraction with Ultra-Performance Liquid Chromatography–Tandem Mass Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Enzymatic Performance of Immobilized Pseudomonas fluorescens Lipase on ZIF-8@ZIF-67 and Its Application to the Synthesis of Neryl Acetate with Transesterification Reaction

School of Chemistry and Chemical Engineering, Henan University of Science and Technology, Luoyang 471023, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(12), 2922; https://doi.org/10.3390/molecules29122922
Submission received: 28 May 2024 / Revised: 16 June 2024 / Accepted: 18 June 2024 / Published: 19 June 2024

Abstract

:
In this study, hybrid skeleton material ZIF-8@ZIF-67 was synthesized by the epitaxial growth method and then was utilized as a carrier for encapsulating Pseudomonas fluorescens lipase (PFL) through the co-precipitation method, resulting in the preparation of immobilized lipase (PFL@ZIF-8@ZIF-67). Subsequently, it was further treated with glutaraldehyde to improve protein immobilization yield. Under optimal immobilization conditions, the specific hydrolytic activity of PFL@ZIF-8@ZIF-67 was 20.4 times higher than that of the free PFL. The prepared biocatalyst was characterized and analyzed by scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier transform infrared (FT-IR). Additionally, the thermal stability of PFL@ZIF-8@ZIF-67 at 50 °C was significantly improved compared to the free PFL. After 7 weeks at room temperature, PFL@ZIF-8@ZIF-67 retained 78% of the transesterification activity, while the free enzyme was only 29%. Finally, PFL@ZIF-8@ZIF-67 was applied to the neryl acetate preparation in a solvent-free system, and the yield of neryl acetate reached 99% after 3 h of reaction. After 10 repetitions, the yields of neryl acetate catalyzed by PFL@ZIF-8@ZIF-67 and the free PFL were 80% and 43%, respectively.

1. Introduction

Lipase (EC 3.1.1.3), an important industrial enzyme that catalyzes reactions such as esterification, transesterification, enantioselective hydrolysis, and kinetic resolution of racemates, is widely used in the food, energy, detergent, pharmaceutical, leather, textile, cosmetics, and paper industries [1,2,3,4,5]. Lipase is a very special enzyme, having a peculiar mechanism of action called interfacial activation [6]. The lipase activity center is flexible and covered by a relatively conserved polypeptide chain structure called a ‘lid’. The lipase lid has a hydrophilic external surface and a hydrophobic internal surface. In homogenous media, a large percentage of most lipase molecules have their active centers covered by the lid, which may isolate them from the reaction medium (closed form) [7]. In the presence of a hydrophobic surface, the enzyme becomes adsorbed on it, the lid structure will be opened, and its active center will be exposed. In aqueous and homogeneous systems, this open form is unstable [8]. The relatively high cost of lipase as a catalyst, its unstable structure, and the difficulty of recovery have limited its further use in industrial applications [9]. To overcome the many drawbacks of free lipase, enzyme immobilization techniques have been investigated. In comparison to the free enzyme, the immobilized enzyme has several advantages, such as easier separation and recycling, higher activity, more stable enzymes, more resistance to various environmental changes, non-toxicity and biocompatibility, or coupling with enzyme purification, making it suitable for more complex reaction environments and a variety of reactor configurations [3,10,11,12]. However, some lipases react with hydrophobic matrixes or other lipase molecules during the immobilization process, resulting in lipase inactivation. In general, the immobilized enzyme properties are affected by carrier materials, immobilization techniques, and enzyme properties [3,13]. The selections of the optimum carrier and the immobilization technique are very important for enzyme immobilization success.
Metal organic frameworks (MOFs) are excellent carriers for immobilized lipases because of their porosity, high specific surface areas, modifiable structure, adjustable pore channels, designable frameworks, and easy functionalization [14,15]. These studies have shown that embedding lipases in MOFs to construct immobilized enzyme systems is more conducive to improving the inherent fragility of lipases and, at the same time, to endowing MOFs with new biological functions [16]. To further improve the performance of single MOFs, more and more researchers have synthesized MOF hybrid frameworks (MOF@MOF) with controlled components, complex morphology, and structure [17]. MOF@MOF is a mixed framework formed by various chemical interactions between different MOFs that often has a core shell or a layered heterostructure. Due to the good synergistic effect among different MOFs materials, these hybrid frameworks often exhibit multifunctional properties and are widely used in wastewater treatment, chemical catalysis, and sensor detection [18,19,20]. Zeolitic imidazolate frameworks (ZIFs) are the most commonly studied materials for MOF@MOF. Research in biochemistry is mainly focused on biosensor detection. ** of S elements on the ZIF-8@ZIF-67 carrier was investigated to determine the distribution of lipase on the carrier. As shown in Figure 2c,d, the element S was evenly distributed on the surface of the carrier, demonstrating that the lipase was immobilized on the ZIF-8@ZIF-67 carrier.

2.2.2. XRD Analysis

The XRD patterns of PFL@ZIF-8@ZIF-67 and ZIF-8@ZIF-67 are shown in Figure 3. Both crystals had strong peaks at 2θ of 10.98°, 14.83°, 16.75°, 18.20°, and 21.87°, consistent with the previously reported ZIF-8@ZIF-67 diffraction [37]. This behavior indicated that the immobilization of the enzyme did not affect the crystal structure of ZIF-8@ZIF-67.

2.2.3. FT-IR Analysis Analysis

Figure 4 shows the IR spectra of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67, in which the representative Zn-N stretching band of ZIF-8@ZIF-67 is observed at approximately 418 cm−1. The strong adsorption bands at 758 cm−1 and 692. 4 cm−1 belonged to the out-of-plane bending of the Hmim ring, while those at 952.8 cm−1 and 1307.7 cm−1 are from in-plane bending. The characteristic peaks of imidazole ligands were strong in these regions, which indicated that organic ligands participated in the reaction. In addition, two weak peaks at 2930 cm−1 and 3124 cm−1 can be attributed to the aliphatic and aromatic CeH stretching of Hmim, respectively, which was consistent with the structure of ZIF-8@ZIF-67 reported in the literature, demonstrating the successful synthesis of ZIF-8@ZIF-67 in this experiment. After the immobilization of lipase, the IR spectrum of PFL@ZIF-8@ZIF-67 showed the characteristic amide I band of lipase at 1658.7 cm−1, corresponding to the NeH bending mode, confirming that lipase was successfully immobilized on ZIF-8@ZIF-67 [38,39].

2.2.4. Michaelis–Menten Kinetic Parameters of PFL@ZIF-8@ZIF-67

The kinetic constants of the enzymes, including Km and Vmax of PFL@ZIF-8@ZIF-67 and the free PFL, were obtained by linear regression of different substrate concentrations and initial reaction velocity catalyzed by the enzymes (Figure 5). The Vmax value of PFL@ZIF-8@ZIF-67 was 4.13 Mm/min, which was greater than the free PFL (2.24 mM/min). The significant increase in Vmax could indicate that the lipase immobilized on ZIF-8@ZIF-67 has a higher catalytic efficiency than the free enzyme, which is related to the activation of the lipase [40]. The Km value of the immobilized lipase (0.36 mM) was lower than that of the free lipase (0.76 mM), indicating that the affinity of the enzyme to the substrate was improved after immobilization. Some studies indicated that the decrease in Km was related to the hydrophobic support and the open conformation of lipase [41,42].

2.3. Application of PFL@ZIF-8@ZIF-67 in the Synthesis of Neryl Acetate

2.3.1. Effect of Immobilized Lipase Amount

From the influence of the amount of immobilized enzyme in Figure 6a, it can be seen that, at the same reaction time, with the increase in the immobilized enzyme amount, the yield gradually increased, especially in the initial stage of the reaction (before 30 min). At the same time, it could be seen from the figure that the yield of neryl acetate obviously increased at the initial stage of the reaction. After 2 h of reaction, the growth tends to be stable. In a certain range of enzyme concentrations, the higher the enzyme concentration, the higher the probability of enzyme contact with the substrate, and the faster the catalytic rate. The amount of enzyme is related not only to the reaction speed but also has an important impact on the production cost of products. Taking into account the factors of yield and economic benefit, 24 mg/mL of immobilized lipase was selected as the catalyst addition.

2.3.2. Effect of Reaction Temperature

Temperature is an important parameter that affects enzyme-catalyzed reactions, and choosing the appropriate reaction temperature not only helps to prolong the life of the enzyme but also increases the rate of the reaction [43]. From the effect of the reaction temperature (Figure 6b), it could be seen that the yield of neryl acetate increases with increasing temperature in the experimental temperature range studied. With increasing temperature, the initial reaction speed increased and then decreased, while the yield of neryl acetate continuously increased and finally stabilized. The initial reaction rate and yield were highest at 50 °C. When the temperature reached 60 °C, the initial rate decreased, and the final conversion rate was similar to that at 50 °C. The results indicated that the optimal reaction temperature for the immobilized enzyme to catalyze the reaction was 50 °C. Compared to the optimal reaction temperature of 40 °C for the free enzyme [27], it increased by 10 °C. This also indicated that the thermal stability of the immobilized enzyme was greatly improved compared to that of the free enzyme. This is because the rigidity of the enzyme increases after it is immobilized on the carrier. To make it more active and combine with the substrate molecules, it is necessary to increase the temperature to increase its flexibility and improve the contact of the enzyme with the substrate. Therefore, the catalytic reaction temperature of immobilized lipase is higher than that of the free lipase.

2.3.3. Effect of Rotating Speed

From the effect of the reaction speed as shown in Figure 6c, it can be seen that at the same reaction time, the yield increases with increasing reaction speed. Also, it can be seen that when the reaction speed is greater than 200 r/min, the yield growth rate tends to be basically stable, which indicates that when the reaction speed reaches 200 r/min, the external diffusion restriction in the reaction system is largely eliminated. The immobilized enzyme reaction system was a heterogeneous reaction system, so the combination of enzyme and substrate molecules was restricted by external diffusion, which affects the catalytic efficiency of the enzyme in the organic phase [44]. The influence of external diffusion could be reduced or even eliminated by studying the influence of the shaking table speed. The literature indicated that rotation speed increases the diffusion of the reactants to the interfacial area of the enzyme and, thus, their availability for the reaction by reducing external mass transfer limitations [45,46]. Considering that the increase in product yield was not obvious at higher speeds, 200 r/min was chosen as the reaction speed to reduce the energy consumption as much as possible without affecting the mass transfer effect. After the above optimization, the yield of neryl acetate could reach over 99% after 3 h of reaction. This was a significant improvement on our previous study [27], and although the final yield of neryl acetate was almost the same, the reaction time was greatly reduced. In contrast to other literature, Zong et al. [47] successfully produced neryl acetate by biosynthesis using recombinant Escherichia coli. Although the maximum production of neryl acetate reached 11.712 ± 0.653 mg/L by increasing pyruvate, the process of synthesizing neryl acetate by this method was too complicated, and large-scale industrial production was difficult. Therefore, the synthesis of neryl acetate using PFL@ZIF-8@ZIF-67 has distinct advantages, which may be helpful for the efficient production of neryl acetate in the future.

2.4. Stability of PFL@ZIF-8@ZIF-67

2.4.1. Reusability of PFL@ZIF-8@ZIF-67

The reusability of enzymes is an important evaluating indicator in practical applications. Figure 7a shows the yield of neryl acetate with immobilized enzyme and the free enzyme after 10 cycles. It was observed that the yield decreased after each cycle. The lower enzymatic activity was attributed to protein denaturation and leakage of protein from the carrier. After 10 reuses, the yield catalyzed by the free PFL was only 43%, whereas that by PFL@ZIF-8@ZIF-67 was 80%. The results illustrated that the immobilized enzyme had higher reusability than the original PFL. The improved recyclability of PFL@ZIF-8@ZIF-67 was mainly caused by the prevention of enzyme desorption, since the MOF@MOF framework and the lipase molecule formed a strong interaction [48]. These data confirmed that the minimal decrease in activity of immobilized lipases was absent of enzyme leakage, and the enzyme retained its conformation during 10 repetitive cycles. Such a decrease in reusability can be attributed to the long period of time that the enzyme was dipped in the reactants, especially the product (neryl acetate). The accumulation of these components inside the biocatalyst particle can produce enzyme inhibition/inactivation. The impact of this problem can be solved by applying ultrasound [49,50] or using a solvent with moderate polarity that is able to solubilize efficiently all the reactants [51]. Also, the use of very hydrophobic supports to immobilize the lipase could be chosen in order to avoid the adsorption of hydrophilic compounds [52].

2.4.2. Storage Stability of PFL@ZIF-8@ZIF-67

Generally, the enzymatic activity decreased gradually with the storage time. The storage stabilities of PFL@ZIF-8@ZIF-67 and the free PFL were examined in the closed conical flask at room temperature for 49 days with a 7-day interval. The results are shown in Figure 7b. After seven weeks of storage, PFL@ZIF-8@ZIF-67 and the free lipase displayed that the residual transesterification activities were 79% and 28%, respectively. The increased storage stability might be due to the protective effect provided by MOF@MOF encapsulation around enzyme molecules, which resulted in minimal distortion caused by environment on the active sites of the enzyme [53].

2.4.3. Thermal Stability of PFL@ZIF-8@ZIF-67

Thermal stability is very important for an immobilization method. The free lipase and PFL@ZIF-8@ZIF-67 were incubated at different temperatures for 6 h and lipase activity was measured, as shown in Figure 7c. When the temperature was less than 40 °C, the transesterification activities of the free enzyme and the immobilized enzyme did not decrease. When the temperature was higher than 40 °C, the transesterification activity of PFL@ZIF-8@ZIF-67 decreased, but the degree of decrease was lower than that of the free PFL. At 60 °C, the residual activities of PFL@ZIF-8@ZIF-67 and the free PFL retained 95% and 72%, respectively. The interactions between lipase and MOF@MOF during the immobilization process can be explained by the enzyme’s protection of its tertiary structure [54].

3. Materials and Methods

3.1. Experimental Material

The lipase powder used in the experiments was derived from Pseudomonas fluorescens and purchased from Amano Enzyme Ltd. (Nagoya, Japan). Bovine serum albumin, nerol, p-nitrophenol (p-NP), and p-nitrophenol butyrate (p-NPB) were obtained from Sigma-Aldrich (St. Louis, MO, USA). Standard neryl acetate was supplied by Tokyo Chemical Industry Co. Ltd. (Tokyo, Japan). Other chemicals and solvents are commercially available analytical grade reagents. All the chemicals and enzymes used have not been further treated.

3.2. Preparation of PFL@ZIF-8@ZIF-67

PFL@ZIF-8@ZIF-67 immobilized lipase was prepared by the co-precipitation method, and the effects of various factors on PFL@ZIF-8@ZIF-67 during the preparation process were investigated. The specific experimental steps were as follows: zinc nitrate hexahydrate solution (0.8 mL, 0.5 M) and 2-methylimidazole solution (8 mL, 1.25 M) were ultrasonicated (40 kHz, 5 min) to make them well dispersed; the two solutions were mixed, and the reaction was stirred at room temperature (500 r/min, 1 h), and the white precipitates were collected by centrifugation (8500 r/min, 10 min) and washed with deionized water for 3 times. Then, it was dried in a vacuum drying oven to constant weight, and finally, ZIF-8 was prepared.
ZIF-8 suspension was prepared by adding 90 mg of the above prepared ZIF-8 to 8.8 mL of deionized water and ultrasonication (40 kHz, 5 min) to disperse it; PFL solution (4 mL, 6 mg/mL), cobalt nitrate hexahydrate solution (0.8 mL, 0.3 M), 2-methylimidazole solution (8 mL, 1.75 M), and NaCl solution (8 mL, 0.3 M) were sonicated (40 kHz, 5 min) to make them well dispersed. The above 4 solutions were added simultaneously to the ZIF-8 suspension, and the reaction was stirred at room temperature (500 r/min, 1 h), and glutaraldehyde solution at a concentration of 4% was added slowly dropwise, and the stirring was continued for 1 h. The above mixture was centrifuged, and the purple precipitate was collected and washed with deionized water 3 times until the protein content of the washing solution was zero; finally, the purple precipitate was lyophilized to a constant weight, and PFL@ZIF-8@ZIF-67 immobilized lipase was obtained.

3.3. Determination of Protein and Enzyme Activity

The protein content was determined using the Coomassie brilliant blue method [55], and bovine serum albumin (BSA) was used as the standard protein. Protein immobilization yield is defined as the amount of protein immobilized on the support as a percentage of the total amount of protein. Using p-NPB as substrate, the hydrolysis activity of the enzyme was determined by colorimetry. Dissolve 0.0195 g of p-NPB in 5 mL of acetone, add 1 mL of Triton-100, and make up to 25 mL with phosphate buffer (pH = 7.5, 0.2 M); 1 mL of the p-NPB solution and 14 mL of the phosphate buffer solution were measured, mixed in an Erlenmeyer flask, and pre-heated for 5 min at 40 °C; 20 mg of immobilized PFL or 3 mg of the free PEL was added to the reaction solution, and the reaction was carried out at 40 °C and 200 r/min for 5 min. The reaction was terminated by the addition of 1 M NaOH (5 mL) followed by centrifugation at 10,000 r/min for 10 min. Absorbance was recorded at 410 nm in a UV spectrophotometer (UV-6100, MAPADA, Shanghai, China) and compared with the standard curve of p-NP (end product of the enzymatic reaction). An enzyme activity unit is defined as the amount of p-NP produced per milligram of enzyme per hour. A specific activity is defined as the enzyme activity per milligram of protein.

3.4. Application of PFL@ZIF-8@ZIF-67 in Synthesis of Neryl Acetate

The transesterification activity of PFL@ZIF-8@ZIF-67 was investigated by using nerol and vinyl acetate to synthesize neryl acetate, and its reaction conditions were optimized. The specific experimental steps were as follows: in a closed Erlenmeyer flask, 2 mL of vinyl acetate was added, at which time the vinyl acetate acted as both reactant and solvent, and the concentration of nerol in the reaction system was ensured to be 100 mmol/L. The solution to be reacted was first preheated at 50 °C and 200 r/min for 10 min, then 28 mg of PFL@ZIF-8@ZIF-67 was added. Starting from the addition of PFL@ZIF-8@ZIF-67, 2 μL was taken out at intervals, and the content of neryl acetate was determined using gas chromatography. Ester exchange activity was defined as the amount of enzyme required to catalyze the production of 1 µmol neryl acetate per a minute.

3.5. Analysis Method of Neryl Acetate

The analysis of neryl acetate were performed using a GC apparatus that was equipped with a flame ionization detector and an SGE AC10 stainless steel column (Fuli, Taizhou, China). The area of the internal standard method was adopted, and hexadecane was used as the standard substance. The following specific test conditions were used: the N2 flow rate was 44 mL/min; the H2 flow rate was 40 mL/min; the air flow rate was 400 mL/min; the tail blowing rate was 25 mL/min; the split ratio was 10:1; the column temperature was kept at 165 °C; the injector temperature was 280 °C; and the detector temperature was 280 °C. The initial rate of the reaction is defined as the amount of neryl acetate generated per unit time and per unit volume.

3.6. Characterization

The morphologies of PFL@ZIF-8@ZIF-67 were observed by scanning electron microscopy (SEM, SU-1510, Hitachi, Tokyo, Japan). The Fourier transform infrared (FT-IR) spectrum was scanned from 400 to 4000 cm−1 on a Nexus 670 instrument. X-ray diffraction (XRD) was carried out on an Ultima IV instrument (Tokyo, Japan).

3.7. Kinetics Analysis

The kinetics parameters of lipase and immobilized lipase samples were obtained by determining the initial reaction rates with p-NPB. The Km value and Vmax were calculated according to the Michaelis–Menten equation.

3.8. Reusability

The transesterification of nerol was used as a model reaction to investigate the reaction stability of lipase. First, 100 mmol/L of nerol was weighed in a 50 mL volumetric flask; then, the volume was fixed with vinyl acetate, and 3 mL was taken, respectively, in a stoppered flask. The free lipase (insoluble in vinyl acetate) or the immobilized lipase was added to the reaction at 50 °C, 200 r/min; the reaction time was 3 h, and the initial speed was measured by taking the solution 5 min before the reaction. After each reaction, the reaction solution was filtered out and the lipase was collected and reused.

3.9. Thermal Stability

A certain amount of the free PFL and PFL@ZIF-8@ZIF-67 was added to vinyl acetate (2 mL) preheated at different temperatures for 6 h; then, the solution was discarded and the preheated reaction solution (100 mmol/L nerol dissolved in vinyl acetate, 3 mL) was added to investigate the transesterification activity of the enzyme after heat treatment.

3.10. Storage Stability

The free PFL and PFL@ZIF-8@ZIF-67 were stored at room temperature in the conical flasks with a cover. The storage stability of the enzymes was investigated by taking a certain amount of enzyme weekly using a transesterification reaction system (100 mmol/L nerol dissolved in vinyl acetate, 3 mL) for the preparation of neryl acetate.

4. Conclusions

In conclusion, a new strategy was devised on MOF@MOF for the enhanced enzymatic performance of immobilized lipase. PFL@ZIF-8@ZIF-67 was successfully prepared and exhibited a significant increase in the specific activity. The specific activity of the immobilized enzyme was 20.4 times higher than that of the free enzyme under the optimized conditions. Meanwhile, PFL@ZIF-8@ZIF-67 can be effectively applied in the production of neryl acetate. The yield of neryl acetate reached 99% after 3 h of reaction and could still be maintained at 80% even after 10 cycles. More importantly, PFL@ZIF-8@ZIF-67 displayed better storage stability and thermal stability than the free lipase. This work provides a new outlook for biotechnological importance by immobilizing lipase on the ZIF-8@ZIF-67 and opens the door for its use in the industrial field.

Supplementary Materials

The following supporting information can be downloaded at https://mdpi.longhoe.net/article/10.3390/molecules29122922/s1, Figure S1: Effect of immobilization time on the structure of immobilized lipase: (a) 2 h; (b) 2.5 h.

Author Contributions

W.S. and J.X. conceived the study; Q.W., S.C., S.L. and Y.Z. performed the experiments and analyzed the data; Q.W., X.P. and W.S. wrote the initial draft of the manuscript; J.X. supervised the projected; J.X. and H.X. revised the manuscript critically. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the College Students’ Innovation and Entrepreneurship Training Program (202310464042 and 2023173).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Raw data are available on request.

Acknowledgments

We are grateful to the Analytical and Testing Center of the College of Chemistry and Chemical Engineering of Henan University of Science and Technology for supporting the analysis equipment.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, J.; Pan, Q.S.; Qian, X.K.; Zhou, X.L.; Wang, Y.J.; He, R.J.; Wang, L.T.; Li, Y.R.; Huo, H.; Sun, C.G.; et al. Discovery of triterpenoids as potent dual inhibitors of pancreatic lipase and human carboxylesterase 1. J. Enzym. Inhib. Med. Chem. 2022, 37, 629–640. [Google Scholar] [CrossRef] [PubMed]
  2. Ali, S.; Khan, S.A.; Hamayun, M.; Lee, I.-J. The recent advances in the utility of microbial lipases: A review. Microorganisms 2023, 11, 510. [Google Scholar] [CrossRef] [PubMed]
  3. Raffee, F.; Rezaee, M. Different strategies for the lipase immobilization on the chitosan based supports and their applications. Int. J. Biol. Macromol. 2021, 179, 170–195. [Google Scholar] [CrossRef] [PubMed]
  4. Facin, B.R.; Melchiors, M.S.; Valerio, A.; Oliveira, J.V.; de Oliveira, D. Driving immobilized lipases as biocatalysts: 10 years state of the art and future prospects. Ind. Eng. Chem. Res. 2019, 58, 5358–5378. [Google Scholar] [CrossRef]
  5. Salgado, C.A.; dos Santos, C.I.A.; Vanetti, M.C.D. Microbial lipases: Propitious biocatalysts for the food industry. Food Biosci. 2022, 45, 101509. [Google Scholar] [CrossRef]
  6. Verger, R. ‘Interfacial activation’ of lipases: Facts and artifacts. Trends Biotechnol. 1997, 15, 32–38. [Google Scholar] [CrossRef]
  7. Schmid, R.R.; Verger, D. Lipases: Interfacial enzymes with attractive applications. Angew. Chem. Int. Ed. 1998, 37, 1608–1633. [Google Scholar] [CrossRef]
  8. Kim, K.K.; Song, H.K.; Shin, D.H.; Hwang, K.Y.; Suh, S.W. The crystal structure of a triacylglycerol lipase from Pseudomonas cepacia reveals a highly open conformation in the absence of a bound inhibitor. Structure 1997, 5, 173–185. [Google Scholar] [CrossRef] [PubMed]
  9. Cao, Y.P.; Zhi, G.Y.; Han, L.; Chen, Q.; Zhang, D.H. Biosynthesis of benzyl cinnamate using an efficient immobilized lipase entrapped in nano-molecular cages. Food Chem. 2021, 364, 130428. [Google Scholar] [CrossRef]
  10. Remonatto, D.; Miotti, R.H.; Monti, R.; Bassan, J.C.; de Paula, A.V. Applications of immobilized lipases in enzymatic reactors: A review. Process Biochem. 2022, 114, 1–20. [Google Scholar] [CrossRef]
  11. Arana-Pena, 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] [PubMed]
  12. Holla, V.; Karkeszova, K.; Antosova, M.; Polakovic, M. Transglycosylation properties of a Kluyveromyces lactis enzyme preparation: Production of tyrosol β-fructoside using free and immobilized enzyme. Process Biochem. 2021, 110, 168–175. [Google Scholar] [CrossRef]
  13. 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]
  14. Jia, Z.R.; Zhang, X.Y.; Gu, Z.; Wu, G.L. MOF-derived Ni-Co bimetal/porous carbon composites as electromagnetic wave absorber. Adv. Compos. Hybrid. Mater. 2022, 6, 28. [Google Scholar] [CrossRef]
  15. Ren, M.Z.; Zhang, Y.J.; Yu, L.L.; Qu, L.B.; Li, Z.H.; Zhang, L. A Co-based MOF as nanozyme with enhanced oxidase-like activity for highly sensitive and selective colorimetric differentiation of aminophenol isomers. Talanta 2022, 255, 124291. [Google Scholar] [CrossRef] [PubMed]
  16. Lian, X.Z.; Fang, Y.; Joseph, E.; Wang, Q.; Li, J.L.; Banerjee, S.; Lollar, C.; Wang, X.; Zhou, H.C. Enzyme-MOF (metal-organic framework) composites. Chem. Soc. Rev. 2017, 46, 3386–3401. [Google Scholar] [CrossRef] [PubMed]
  17. Hong, D.H.; Shim, H.S.; Ha, J.; Moon, H.R. MOF-on-MOF Architectures: Applications in Separation, Catalysis, and Sensing. Bull. Korean Chem. Soc. 2021, 42, 956–969. [Google Scholar] [CrossRef]
  18. Mao, J.W.; Meng, Q.; Xu, Z.H.; Xu, L.S.; Fan, Z.; Zhang, G.L. MOF-on-MOF heterojunction-derived Co3O4-CuCo2O4 microflowers for low-temperature catalytic oxidation. Chem. Commun. 2022, 58, 13600–13603. [Google Scholar] [CrossRef] [PubMed]
  19. Li, J.; Yu, X.M.; Xue, W.J.; Nie, L.; Huang, H.L.; Zhong, C.L. Engineering the direct Z-scheme systems over lattice intergrown of MOF-on-MOF for selective CO2 photoreduction to CO. AIChE J. 2022, 69, e17906. [Google Scholar] [CrossRef]
  20. Yao, M.S.; **u, J.W.; Huang, Q.Q.; Li, W.H.; Wu, W.W.; Wu, A.Q.; Cao, L.A.; Deng, W.H.; Wang, G.E.; Xu, G. Van der waals heterostructured MOF-on-MOF thin films: Cascading functionality to realize advanced chemiresistive sensing. Angew. Chem. Int. Ed. 2019, 58, 4915–14919. [Google Scholar]
  21. **ao, L.L.; Zheng, S.Y.; Yang, K.X.; Duan, J.X.; Jiang, J. The construction of CoP nanoparticles coated with carbon layers derived from core-shell bimetallic MOF for electrochemical detection of dopamine. Microchem. J. 2021, 168, 106432. [Google Scholar] [CrossRef]
  22. Cao, Q.; **ao, Y.S.; Liu, N.; Huang, R.; Ye, C.; Huang, C.H.; Liu, H.; Han, G.; Wu, L.D. Synthesis of Yolk/Shell Heterostructures MOF@MOF as Biomimetic Sensing Platform for Catechol Detection. Sens. Actuators B 2020, 370, 132503. [Google Scholar] [CrossRef]
  23. Zeferino, R.C.F.; Piaia, V.A.A.; Orso, V.T.; Pinheiro, V.M.; Zanetti, M.; Colpani, G.L.; Padoin, N.; Soares, C.; Fiori, M.A.; Riella, H.G. Neryl acetate synthesis from nerol esterification with acetic anhydride by heterogeneous catalysis using ion exchange resin. J. Ind. Eng. Chem. 2022, 105, 121–131. [Google Scholar] [CrossRef]
  24. Singh, A.P.; Arya, H.; Singh, V.; Kumar, P.; Gautam, H.K. Identification of natural inhibitors to inhibit C. acnes lipase through docking and simulation studies. J. Mol. Model. 2022, 28, 281. [Google Scholar] [CrossRef] [PubMed]
  25. Parreira, L.A.; Azevedo, A.F.; Menini, L.; Gusevskaya, E.V. Functionalization of the naturally occurring linalool and nerol by the palladium catalyzed oxidation of their trisubstituted olefinic bonds. J. Mol. Catal. A Chem. 2017, 426, 429–434. [Google Scholar] [CrossRef]
  26. Lozano, P.; Bernal, J.M.; Navarro, A. A clean enzymatic process for producing flavour esters by direct esterification in switchable ionic liquid/solid phases. Green Chem. 2012, 14, 3026–3033. [Google Scholar] [CrossRef]
  27. Sun, W.Y.; **ong, J.; Xu, H.H.; Ma, M.Y.; Hu, Y.Y. Synthesis of neryl acetate by free lipase-catalyzed transesterification in organic solvents and its kinetics. Food Sci. Technol. 2022, 42, e26522. [Google Scholar] [CrossRef]
  28. Xue, W.; Yang, G.; Bi, S.; Zhang, J.Y.; Hou, Z.L. Construction of caterpillar-like hierarchically structured Co/MnO/CNTs derived from MnO2/ZIF-8@ZIF-67 for electromagnetic wave absorption. Carbon 2021, 173, 521–527. [Google Scholar] [CrossRef]
  29. Pan, Y.; Sun, K.; Liu, S.J.; Cao, X.; Wu, K.L.; Cheong, W.C.; Chen, Z.; Wang, Y.; Li, Y.; Liu, Y.Q.; et al. Core-shell ZIF-8@ZIF-67 derived CoP nanoparticles-embedded N-doped carbon nanotube hollow polyhedron for efficient over-all water splitting. J. Am. Chem. Soc. 2018, 140, 2610–2618. [Google Scholar] [CrossRef]
  30. Lu, L.D.; Zhu, L.Q.; Zhu, G.X.; Dong, M.L.; Zeng, Z.M. ZIF-8/lipase coated tapered optical fiber biosensor for the detection of triacylglycerides. IEEE Sens. J. 2020, 20, 14173–14180. [Google Scholar] [CrossRef]
  31. Vaidya, L.B.; Nadar, S.S.; Rathod, V.K. Entrapment of surfactant modified lipase within zeolitic imidazolate framework (ZIF)-8. Int. J. Biol. Macromol. 2020, 146, 678–686. [Google Scholar] [CrossRef] [PubMed]
  32. Adnan, M.; Li, K.; Wang, J.H.; Xu, L.; Yan, Y.J. Hierarchical ZIF-8 toward immobilizing Burkholderia cepacia lipase for application in biodiesel preparation. Int. J. Mol. Sci. 2018, 19, 1424. [Google Scholar] [CrossRef] [PubMed]
  33. Pu, S.J.; Zhang, X.; Yang, C.L.; Naseer, S.; Zhang, X.T.; Ouyang, J.; Li, D.L.; Yang, J.F. The effects of NaCl on enzyme encapsulation by zeolitic imidazolate frameworks-8. Enzym. Microb. Technol. 2019, 122, 1–6. [Google Scholar] [CrossRef] [PubMed]
  34. Barbosa, O.; Ortiz, C.; Berenguer-Murcia, A.; Torres, R.; Rodrigues, R.C.; Fernandez-Lafuente, R. Glutaraldehyde in bio-catalysts design: A useful crosslinker and a versatile tool in enzyme immobilization. RSC Adv. 2013, 4, 1583–1600. [Google Scholar] [CrossRef]
  35. Kohsari, I.; Shariatinia, Z.; Pourmortazavi, S.M. Antibacterial electrospun chitosan-polyethylene oxide nanocomposite mats containing ZIF-8 nanoparticles. Int. J. Biol. Macromol. 2016, 91, 778–788. [Google Scholar] [CrossRef] [PubMed]
  36. Tasarin, S.; Panawong, C.; Sumranjit, J.; Budsombat, S. Enhancement of proton conductivity of crosslinked poly(vinyl alcohol) through introduction of zeolitic imidazolate framework-8 and imidazole. Int. J. Hydrog. Energy 2021, 46, 36969–36981. [Google Scholar] [CrossRef]
  37. Lv, B.Z.; Qi, W.J.; Luo, M.S.; Liu, Q.L.; Guo, L. Fischer-tropsch synthesis: ZIF-8@ZIF-67-derived cobalt nanoparticle-embedded nanocage catalysts. Ind. Eng. Chem. Res. 2020, 59, 12352–12359. [Google Scholar]
  38. Li, Y.; Zhou, K.; He, M.; Yao, J.F. Synthesis of ZIF-8 and ZIF-67 using mixed-base and their dye adsorption. Microporous Mesoporous Mater. 2016, 234, 287–292. [Google Scholar] [CrossRef]
  39. Zhao, Y.J.; Ding, L.; Wang, X.M.; Yang, X.M.; He, J.B.; Yang, B.J.; Wang, B.N.; Zhang, D.W.; Li, Z.W. Yolk-shell ZIF-8@ZIF-67 derived Co3O4@NiCo2O4 catalysts with effective electrochemical properties for Li-O2 batteries. J. Alloys Compd. 2021, 861, 157945. [Google Scholar] [CrossRef]
  40. Guo, Z.; Bai, S.; Sun, Y. Preparation and characterization of immobilized lipase on magnetic hydrophobic microspheres. Enzym. Microb. Technol. 2003, 32, 776–782. [Google Scholar] [CrossRef]
  41. Suo, H.B.; Xu, L.L.; Xu, C.; Qiu, X.; Huang, H.; Hu, Y. Enhanced catalytic performance of lipase covalently bonded on ionic liquids modified magnetic alginate composites. J. Colloid. Interface Sci. 2019, 553, 494–502. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, C.Y.; Dong, X.Y.; Guo, Z.; Sun, Y. Remarkably enhanced activity and substrate affinity of lipase covalently bonded on zwitterionic polymer-grafted silicananoparticles. J. Colloid. Interface Sci. 2018, 519, 145–153. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, X.; Wang, X.; Cong, F.; Xu, Y.; Kang, J.; Zhang, Y.; Zhou, M.; **ng, K.; Zhang, G.; Pan, H. Synthesis of cinnamyl acetate catalysed by highly reusable cotton-immobilized Pseudomonas fuorescens lipase. Biocatal. Biotransfor. 2018, 36, 332–339. [Google Scholar] [CrossRef]
  44. De Castro, H.F.; De Oliveira, P.C.; Pereira, E.B. Evaluation of different approaches for lipase catalysed synthesis of citronellyl acetate. Biotech. Lett. 1997, 19, 229–232. [Google Scholar] [CrossRef]
  45. Dossat, V.; Combes, D.; Marty, A. Efficient lipase catalysed production of a lubricant and surfactant formulation using a continuous solvent-free process. J. Biontechnol. 2002, 97, 117–124. [Google Scholar] [CrossRef]
  46. Kumari, A.; Mahapatra, P.; Garlapati, V.; Banerjee, R. Enzymatic transesterification of Jatropha oil. Biotechnol. Biofuels 2009, 2, 1. [Google Scholar] [CrossRef] [PubMed]
  47. Zong, Z.; Zhang, S.P.; Zhen, M.L.; Xu, N. Metabolic engineering of Escherichia coli for the production of neryl acetate. Biochem. Eng. J. 2020, 161, 107704. [Google Scholar] [CrossRef]
  48. He, H.M.; Han, H.B.; Shi, H.; Tian, Y.Y.; Sun, F.X.; Song, Y.; Li, Q.S.; Zhu, G.S. Construction of thermophilic lipase-embedded metal organic frameworks via biomimetic mineralization: A biocatalyst for ester hydrolysis and kinetic resolution. ACS Appl. Mater. Interfaces 2016, 8, 24517–24524. [Google Scholar] [CrossRef]
  49. Martins, A.B.; Schein, M.F.; Friedrich, J.L.R.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Ultrasound-assisted butyl acetate synthesis catalyzed by Novozym 435: Enhanced activity and operational stability. Ultrason. Sonochem. 2013, 20, 1155–1160. [Google Scholar] [CrossRef]
  50. Graebin, N.G.; Martins, A.B.; Lorenzoni, A.S.G.; Garcia-Galan, C.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Immobilization of lipase B from Candida antarctica on porous styrene–divinylbenzene beads improves butyl acetate synthesis. Biotechnol. Prog. 2012, 28, 406–412. [Google Scholar] [CrossRef]
  51. Chiaradia, V.; Valerio, A.; De Oliveira, D.; Araújo, P.H.H.; Sayer, C. Simultaneous single-step immobilization of Candida antarctica lipase B and incorporation of magnetic nanoparticles on poly(urea-urethane) nanoparticles by interfacial miniemulsion polymerization. J. Mol. Catal. B Enzym. 2016, 131, 31–35. [Google Scholar] [CrossRef]
  52. Severac, E.; Galy, O.; Turon, F.; Pantel, C.A.; Condoret, J.S.; Monsan, P.; Marty, A. Selection of CalB immobilization method to be used in continuous oil transesterification: Analysis of the economical impact. Enzym. Microb. Technol. 2011, 48, 61–70. [Google Scholar] [CrossRef] [PubMed]
  53. Kumari, G.; Jayaramulu, K.; Maji, T.K.; Narayana, C. Temperature induced structural transformations and gas adsorption in the zeolitic imidazolate framework ZIF-8: A Raman study. J. Phys. Chem. A 2013, 117, 11006–11012. [Google Scholar] [CrossRef] [PubMed]
  54. Ozyilmaz, E.; Ascioglu, S.; Yilmaz, M. Calix[4]arene tetracarboxylic acid-treated lipase immobilized onto metal-organic framework: Biocatalyst for ester hydrolysis and kinetic resolution. Int. J. Biol. Macromol. 2021, 175, 79–86. [Google Scholar] [CrossRef]
  55. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef]
Figure 1. Optimization of the preparation conditions for PFL@ZIF-8@ZIF-67: (a) amount of ZIF-8; (b) concentration of cobalt nitrate; (c) concentration of 2-methylimidazole; (d) concentration of lipase; (e) immobilization time; (f) concentration of sodium chloride; (g) concentration of glutaraldehyde.
Figure 1. Optimization of the preparation conditions for PFL@ZIF-8@ZIF-67: (a) amount of ZIF-8; (b) concentration of cobalt nitrate; (c) concentration of 2-methylimidazole; (d) concentration of lipase; (e) immobilization time; (f) concentration of sodium chloride; (g) concentration of glutaraldehyde.
Molecules 29 02922 g001aMolecules 29 02922 g001b
Figure 2. SEM analysis of PFL@ZIF-8@ZIF-67: (ac) SEM images; (d) the map** of S elements on a local scale.
Figure 2. SEM analysis of PFL@ZIF-8@ZIF-67: (ac) SEM images; (d) the map** of S elements on a local scale.
Molecules 29 02922 g002
Figure 3. XRD comparison of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Figure 3. XRD comparison of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Molecules 29 02922 g003
Figure 4. FT-IR comparison of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Figure 4. FT-IR comparison of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Molecules 29 02922 g004
Figure 5. Lineweaver–Burk plot of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Figure 5. Lineweaver–Burk plot of ZIF-8@ZIF-67 and PFL@ZIF-8@ZIF-67.
Molecules 29 02922 g005
Figure 6. Effect of reaction conditions on the yield of neryl acetate: (a) immobilized lipase amount; (b) temperature; (c) rotating speed.
Figure 6. Effect of reaction conditions on the yield of neryl acetate: (a) immobilized lipase amount; (b) temperature; (c) rotating speed.
Molecules 29 02922 g006
Figure 7. Stability of reaction conditions on the yield of neryl acetate: (a) reusability; (b) storage stability; (c) thermal stability.
Figure 7. Stability of reaction conditions on the yield of neryl acetate: (a) reusability; (b) storage stability; (c) thermal stability.
Molecules 29 02922 g007
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

Wang, Q.; **ong, J.; Xu, H.; Sun, W.; Pan, X.; Cui, S.; Lv, S.; Zhang, Y. Enhanced Enzymatic Performance of Immobilized Pseudomonas fluorescens Lipase on ZIF-8@ZIF-67 and Its Application to the Synthesis of Neryl Acetate with Transesterification Reaction. Molecules 2024, 29, 2922. https://doi.org/10.3390/molecules29122922

AMA Style

Wang Q, **ong J, Xu H, Sun W, Pan X, Cui S, Lv S, Zhang Y. Enhanced Enzymatic Performance of Immobilized Pseudomonas fluorescens Lipase on ZIF-8@ZIF-67 and Its Application to the Synthesis of Neryl Acetate with Transesterification Reaction. Molecules. 2024; 29(12):2922. https://doi.org/10.3390/molecules29122922

Chicago/Turabian Style

Wang, Qi, Jian **ong, Hanghang Xu, Wenyuan Sun, **aoxu Pan, Shixin Cui, Siting Lv, and Yinling Zhang. 2024. "Enhanced Enzymatic Performance of Immobilized Pseudomonas fluorescens Lipase on ZIF-8@ZIF-67 and Its Application to the Synthesis of Neryl Acetate with Transesterification Reaction" Molecules 29, no. 12: 2922. https://doi.org/10.3390/molecules29122922

Article Metrics

Back to TopTop