Alice
Vetrano
a,
Francesco
Gabriele
a,
Raimondo
Germani
b and
Nicoletta
Spreti
*a
aDepartment of Physical and Chemical Sciences, University of L’Aquila, Via Vetoio – Coppito, I-67100 L’Aquila, Italy. E-mail: nicoletta.spreti@univaq.it
bCEMIN, Centre of Excellence on Nanostructured Innovative Materials, Department of Chemistry, Biology and Biotechnology, University of Perugia, Via Elce di Sotto 8, I-06123 Perugia, Italy
First published on 4th May 2022
Lipase from Candida rugosa has been immobilized in different formulations of calcium alginate beads, prepared by ionotropic gelation, which differ from each other in CaCl2 concentration and hardening time, to investigate the effects of immobilization conditions on enzyme properties. Morphological studies on all hydrated beads, performed by SEM equipped with a Peltier plate, revealed a different internal compactness. Despite this, all types of beads had an immobilization yield of 100% measured with the Bradford method and about 94% evaluated from the residual activity of the preparation solutions; moreover, all entrapped biocatalysts catalyzed the complete hydrolysis of p-nitrophenyl acetate, even after one month of storage in distilled water at 4 °C. When the internal microstructure of the beads was more compact, the rate of hydrolysis of the most hydrophobic p-nitrophenyl dodecanoate was halved, probably due to mass transfer limiting effects. The immobilized lipase had better resistance to temperature inactivation than the free form: enzyme residual activity at 50 °C after a week were approximately 70% and 20% for the immobilized and free forms respectively. An excellent recyclability in water at 25 °C of entrapped enzyme was also found, having residual activity greater than 80% at the tenth reaction cycle. The best bead formulation was then used for the resolution of (R)-1-phenylethanol in aqueous solution starting from racemic (R,S)-1-phenylethyl acetate. The enantioselectivity found (E = 10) was slightly higher but did not differ significantly from that of free lipase towards the same substrate (E = 4).
Lipases (triacylglycerol ester hydrolase, E.C. 3.1.1.3) are ubiquitous enzymes widely employed to catalyze a wide range of enantio- and regioselective reactions such as hydrolysis, esterification, transesterification, aminolysis and ammoniolysis.14–18 Unlike the usual esterases, lipases are able to hydrolyze long-chain acyl glycerols and contain an amphiphilic peptide lid domain covering the active site of the enzyme, which, in the presence of a hydrophobic interface, undergoes a conformational rearrangement allowing the enzyme to switch to the active state, a phenomenon known as interfacial activation.19
Lipase from Candida rugosa (CRL), whose structural features, mechanism of action and catalytic versatility are well known, exists as several isoenzymes, with a high structural homology, but different carbohydrate content, isoelectric point and substrate specificity.20–23 CRL is one of the most used enzymes for biotransformations, but the use of the free form is not convenient, as it is deactivated when exposed to temperatures higher than 50 °C for a long time24 and, above all, the enzyme is not recyclable and its separation from the final products requires many time-consuming steps. For these reasons, several methods have been reported to improve enzyme stability and recyclability: physical adsorption on solid supports,25–29 cross-linking,30–32 covalent binding33–35 and encapsulation on a solid matrix.36–39 Adsorption is the simplest and cheapest method, but the leaching of the enzyme from the support is a significant disadvantage for practical uses, compromising its reusability. Cross-linked enzyme aggregates, employing bifunctional reagents, can be either used as carrier-free macroparticles or immobilized on different supports30,31 or they can be trapped in alginate beads.32 In these cases, the operational stability and reusability of the biocatalyst have been improved and the immobilization protocol can be decisive due to the variety of enzyme structural conformations. Covalent binding to a support generally prevents enzyme leaching from the surface, but this method could have the disadvantage of a possible irreversible deactivation. Reusability, storage stability and tolerance to organic solvents were improved when CRL, chemically modified with a monomer of the metal–organic frameworks was used to the “in situ” synthesis of immobilized CRL composite,33 as well as better lipase activity was achieved as a result of irreversible enzyme immobilization onto a ternary alginate/nanocellulose/montmorillonite composite.34 The covalent immobilization of CRL onto magnetic beads offered important advantages in terms of enzyme reusability, thanks to the magnetically easy recovery of the biocatalyst from the reaction media, and can be applied for synthetic purpose,35 for food applications,40,41 and for biodiesel production.42,43 Moreover, CRL, encapsulated in silica sol gels in the presence of magnetic sporopollenin/Fe3O4 nanoparticles,36,37 β-cyclodextrin-grafted38 and in the presence of N-methylglucamine based calix[4]arene magnetic nanoparticles39 exhibited high thermal stability, reusability and excellent enantioselective capability.
Naturally-derived polymers turn out to be more advantageous in the entrapment of biomolecules than synthetic ones thanks to their intrinsic properties such as biocompatibility, non-toxicity, biodegradability and renewability that make them very attractive supports in many applications in biomedical, pharmaceutical and food sectors.44 Among them, alginate, an anionic polysaccharide derived from brown algae consisting of β-D-mannuronate (M) and α-L-guluronate (G) as monomeric units, is by far the most widely used polymer for immobilization and microencapsulation technologies, thanks to its ability to easily form the desired three-dimensional structures in an aqueous environment by coordinating divalent cations.45 Many alginate-based supports have been developed for enzyme immobilization and for the enhancement of enzyme properties, in terms of operational stability and reusability, together with their biotechnological applications, have recently been reviewed.44 CRL was also entrapped in Ca-alginate beads and, thanks to the strong affinity of the polysaccharide for the enzyme, its thermal stability was improved as the energy barrier of the first deactivation step was higher than that of the free enzyme.24 Nevertheless, the main problem was the CRL leaching from the beads,46 that can be partially overcome by covering their surface with chitosan or silicate.47 Esterification reactions were successfully performed in aqueous media using biphasic alginate beads, consisting of a solid matrix of calcium alginate and hexadecane,48 or in organic solvents when CRL was immobilized in polyvinyl alcohol (PVA), alginate and boric acid beads.49 In the latter case, the biocatalyst was more compatible with both water and organic media, since a certain amount of water molecules was also allowed in dehydrating solvents. The immobilization of CRL in magnetic alginate beads made possible to easily collect and reuse the biocatalyst for 6 cycles, but its activity was lower with respect to the enzyme entrapped in alginate beads.50
In this paper, CRL was immobilized within six different formulations of calcium-alginate beads. The operating conditions, such as the concentration of CaCl2 and the residence time in the hardening solution, have been taken into account to evaluate their effect on: (i) loading efficiency; (ii) biocatalytic hydrolysis toward two substrates with different hydrophobicity p-nitrophenyl acetate (p-NPA) and p-nitrophenyl dodecanoate (p-NPD); (iii) recyclability and thermostability of the immobilized biocatalyst, decisive properties for industrial applications, and (iv) internal structure. Finally, the best formulation was chosen to perform the kinetic resolution of the racemic ester (R,S)-1-phenylethyl acetate.
CaCl2, % (w/v) | Residence time, min | |
---|---|---|
a Sodium alginate solution 5% (w/v) and CRL 2 mg ml−1. | ||
Beads 1 | 2 | 10 |
Beads 2 | 5 | 10 |
Beads 3 | 2 | 30 |
Beads 4 | 5 | 30 |
Beads 5 | 2 | 60 |
Beads 6 | 5 | 60 |
![]() | (1) |
![]() | (2) |
Activity of encapsulated lipase was also determined with a more hydrophobic substrate, p-nitrophenyl dodecanoate (p-NPD), dissolved in CH3CN; 1 ml of a 100 mM stock solution was placed in the reaction vessel containing 9 ml of tert-butyl alcohol as a solvent and the beads containing the enzyme (1.2 ÷ 1.4 g – 4 mg CRL). The reaction was monitored at different times following the appearance of p-NP at 348 nm.
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | (7) |
Sphericity factor (SF), indicating the roundness of the beads, was determined by using the following equation:55,56
![]() | (8) |
The aspect ratio (AR) gives a good description of large bead deformations but is less accurate on smaller ones. AR varies from unity for a sphere to infinity for an elongated particle and was determined using the following equation:
![]() | (9) |
The surface morphology and the internal structure of the hydrate systems was investigated using a scanning electron microscope (SEM) equipped with a Peltier cooling-device MK3 Cool stage Carl Zeiss SUPRA with a working distance of about 8 mm and high voltage of 10 KV. Analyses were done around 0 °C in variable pressure mode (20 Pa) using a BSE detector (Signal A BSD4).
Fig. 1 shows a schematic representation of the synthetic process for preparing the immobilized CRL on Ca-alginate beads.
Once this was established, activity tests were performed with the CRL-containing beads in pure water, rather than in buffer solution. Indeed, notwithstanding the pH of the medium is a critical parameter in enzymatic reactions, since it affects ionization state of enzyme leading to changes in the active site,57 the presence of salts in the reactor vessel could lead to corrosion issues in view of industrial applications, as biofuel production.58
The first catalytic tests were performed with all the bead formulations with p-NPA at a concentration of 10 mM. Fig. 2 shows the conversion percentage of the substrate over time for the six types of beads.
![]() | ||
Fig. 2 Substrate conversion percentages for all types of beads at 30 min (![]() ![]() ![]() ![]() |
The figure clearly shows that, regardless of the bead formulation, all reactions were complete within three hours with an initial reaction rate of 130 ± 10 μM min−1, and only slight differences were observed during the reaction between the different types of beads. Therefore, it was not possible to discriminate them in terms of conversion efficiency if used immediately after their preparation.
In Section 3.2, we proved that there was no leakage of enzyme from the beads during long-term storage, except for Beads 1 and Beads 2, where the loss was still less than or equal to 1%, but we had no information regarding its activity after a long period of storage in water. Therefore, a reaction cycle was performed one month after preparation to evaluate whether the enzyme in the beads was still active. For all types of beads, complete substrate conversion was achieved within four hours indicating that there were no significant changes in the activity of CRL after a month of storage (Fig. S2, ESI†).
To assess the reusability of CRL within Ca-alginate beads, the two extreme formulations, Beads 1 and Beads 6, were chosen to perform ten catalytic cycles, in order to evaluate the influence of both CaCl2 concentration and residence time in the solution itself. These tests were carried out with 10 mM substrate and reactions were followed until to completion. Fig. 3 shows the results of repeated uses obtained after 3 h of reaction, which is the time required to the biocatalyst to complete the first cycle.
![]() | ||
Fig. 3 Residual activity of immobilized CRL in Ca-alginate beads in water after 3 h of reaction: Beads 1 (![]() ![]() |
The data obtained clearly indicated that Beads 1 and Beads 6 were able to convert all the substrate into product for the first four and three cycles respectively and, at the tenth cycle, for both formulations the loss of activity was less than 20%. Despite this slightly decrease of relative activity, the reaction time from 3 h to the first cycle increases up to 7 h to the tenth one to reach the complete hydrolysis of the substrate, due to a halving of initial reaction rate for both formulations (from about 120 μM min−1 to 60 μM min−1). The results obtained so far showed that there were no differences in terms of catalytic efficiency and recyclability between the different formulations.
The use of alginate beads in the immobilization of CRL according to our formulations, compared to other systems, shows numerous advantages, as can be seen from Table 2. The preparation of the support is simple and requires only two steps, and high immobilization efficiency, recyclability and storage stability are achieved. Furthermore, our biocatalyst is active in distilled water, avoiding the buffer which is probably the cause of the observed enzyme leakage46,47 and induces swelling of the beads increasing their fragility (data not shown).
Immobilization method | Preparation step | Biocatalyst | Immobilization yield | Substrate | Recycle times | Residual activity | Operational condition | Storage stability (days) | Residual activity | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
Covalent immobilization onto a solid carrier (MOF) | 2 | Candida rugosa lipase (CRL) type VII | CRL-ZIF-8 8.18%, mCRL-ZIF-8 9.45% | p-Nitrophenyl palmitate | 7 | CRL-ZIF-8 45.8%, mCRL-ZIF-8 52.3% | Isopropanol- PBS pH 7,5 (1![]() ![]() |
7 | CRL-ZIF-8 34.42%, mCRL-ZIF-8 42.18% | 32 |
Covalent immobilization onto a ternary support (CRL-ALG/NC/MMT) | 5 | Candida rugosa lipase (CRL) type VII | 2.9% | Levulinic acid | 9 | 47.5% | Ethanol 50 °C | — | — | 34 |
Covalent immobilization on magnetic beads | 5 | Candida rugosa lipase (CRL) | — | Racemic ibuprofen | 5 | 100% | Cyclohexane 37 °C | — | — | 35 |
Sol–gel encapsulation in presence of magnetic sporopollenin/Fe3O4 nanoparticles | 4 | Candida rugosa lipase (CRL) | Fe-A-Spo-E 36%, Fe-EP-Spo-E 89% | p-Nitrophenyl palmitate | 7 | Fe-A-Spo-E 21%, Fe-EP-Spo-E 63% | Isopropanol – PBS pH 7 37 °C | 50 | Fe-A-Spo-E 45%, Fe-EP-Spo-E 40% | 36 |
Sol–gel encapsulation e with Fe3O4 or Fe3O4-Spo | 5 | Candida rugosa lipase (CRL) type VII | Fe3O4 78.6%, Fe3O4-Spo 76.2% | Racemic naproxen methyl ester | 5 | Fe3O4-E 36%, Fe3O4-Spo 64% | Aqueous phase–organic solvent (PBS pH 7 – isooctane) | 42 | Fe3O4 >70%, Fe3O4-Spo >90% | 37 |
Sol–gel encapsulation with b-cyclodextrin grafted magnetic nanoparticles (CD-APS-NP) | 5 | Candida rugosa lipase (CRL) type VII | 2.2% | p-Nitrophenyl palmitate | 6 | ∼50% | PBS buffer (pH 7.0) 35 °C | — | — | 38 |
Sol–gel Encapsulation in the presence of magnetic calix[4]arene nanoparticles | 4 | Candida rugosa lipase (CRL) | 3.2% | Racemic naproxen methyl ester | 5 | 28% | Aqueous phase–organic solvent (PBS – isooctane) | — | — | 39 |
Covalent immobilization on magnetite coated with nanosilica | 5 | Candida rugosa lipase (CRL) | ∼80% |
n-Butyric acid and 1-butanol (1![]() ![]() |
17 | 50% | n-Heptane at 45 °C | — | — | 59 |
Encapsulation in chitosan nanoparticles | 3 | Candida rugosa lipase (CRL) | — | Olive oil | 7 | 53% | Water–oil (1![]() ![]() |
— | — | 60 |
Entrapment in Ca-alginate beads | 2 | Candida rugosa lipase (CRL) | 35% | p-Nitrophenyl butyrate | 3 | 72% | Tris–HCl buffer (pH 7.2); 30 °C | — | — | 47 |
Covalent immobilization of lipase onto the silica nanoflowers-NH2 | 3 | Candida antarctica lipase | ∼57% | Levulinic acid and ethanol (1![]() ![]() |
8 | 68% | Tert-butyl methyl ether 40 °C | — | — | 61 |
Freeze dried calcium alginate beads | 3 | Candida antarctica lipase B (CALB) | — | p-Nitrophenyl butyrate | 6 | ∼90% | Distilled water 35 °C | 16 | ∼100% | 62 |
Immobilization on magnetic sol–gel hybrid organic-inorganic (Fe3O4 MNPs@TEOS-TSD@ CALB) | 4 | Candida antarctica lipase B (CALB) | 90% | Waste cooking oil | 10 | 16.5% | Methanol; 40 °C | — | — | 63 |
Covalent immobilization on polydopamine functionalized magnetic mesoporous biochar (MPCB-DA) | 6 | Bacillus licheniformis lipase | 46.5% | p-Nitrophenyl palmitate | 10 | 56% | Tris–HCl (50 mM, pH 8.5); 40 °C | 70 | 88% at 25 °C | 64 |
Covalent immobilization on hydroxyapatite/glycyrrhizin/lithium-based metal–organic framework (HA/GL/Li-MOF) nanocomposites | 5 | Thermomyces lanuginosus lipase (TLL) | TLL@HA/GL/Li-MOF 71%, TLL@Li-MOF 68% | p-Nitrophenyl palmitate | 10 | TLL@HA/GL/Li-MOF ∼55%, TLL@Li-MOF ∼45% | Carbonate bufer (100 mM, pH 9.0); 60 °C | 30 | TLL@HA/GL/Li-MOF ∼30%, TLL@Li-MOF ∼20% | 65 |
Entrapment inside Ca-Alginate beads | 2 | Candida rugosa lipase (CRL) type VII | 94.4% | p-Nitrophenyl acetate | 10 | >80% | Distilled water 25 °C | 30 | ∼100% | Current work |
Analyses were initially performed using the stereomicroscope and the results are shown in Fig. 4.
Both Beads 1 and Beads 6 showed fairly uniform dimensions and the size distribution of 25 beads was measured, since the variance coefficient (CV), determined using eqn (7), was 0.9 and 3.7% respectively. Table 3 reports some of the dimensionless shape indicators quantified using eqn (8) and (9) (Section 2.8).
Beads 1 | Beads 6 | |
---|---|---|
Average diameter (Dm) | 3.2 mm | 3.4 mm |
Sphericity factor (SF) | 0.031 | 0.069 |
Aspect ratio (AR) | 1.06 | 1.15 |
Both bead formulations have similar dimensions of about 3 mm, but better spherical shape was obtained with Beads 1, which are those prepared with lower calcium chloride concentration and shorter hardening time, since SF < 0.05 and AR was slightly higher than unity.56 On the other hand, Beads 6, as already visible from the stereomicroscope image, are less spherical; their SF value was very similar to that of alginate particles reported in the literature and prepared with our same experimental conditions, i.e. concentration of alginate and CaCl2, and hardening time.66 Then, the analyses by the scanning electron microscope were performed on both the external surface and the internal structure (by cutting them with a scalpel). Generally, in literature, SEM morphological studies are conducted on dehydrated beads, providing information that often does not correspond to the real systems employed in catalytic tests, which are hydrated. To avoid dehydration of the samples, the application of variable-pressure equipment (VP-SEM) and Peltier cooling-device allows the investigation of wet samples and hydrated systems in SEM.67 Thanks to the aid of this methodology, in this study we were able to investigate the samples in their operating conditions. The SEM images were acquired at different magnifications and the most significant at 70x and 300x are shown in Fig. 5.
![]() | ||
Fig. 5 SEM images at 70x magnification of the external structure and at 70x and 300x magnifications of the internal structure of Beads 1 and Beads 6. |
Although no significant differences were detected from their external surface, the analysis of the cross section highlighted less internal compactness of Beads 1 than that of Beads 6 which seemed denser and more homogeneous. Further SEM analyses were then performed to understand if the cause of this different internal morphology was due to the CaCl2 concentration and/or to the hardening time. For these analyses, Beads 2 (CaCl2 5%, 10 min) and Beads 5 (CaCl2 2%, 60 min) were selected and SEM images, reported in Fig. S3 (ESI†) of both the whole beads and the internal structure highlighted that the internal compactness depends on the gelation time and not on CaCl2 concentration.
Furthermore, the time taken by Beads 1 to completely hydrolyze p-NPD was twice than that required for the reaction with p-NPA. Therefore, to determine if the increase in reaction time was due to the substrate or to the solvent, the reaction of p-NPA in tert-butyl alcohol was performed for comparison purpose (Fig. S5, ESI†). After 48 h, only 58% of conversion was achieved, and this result can be explained by the dehydration of the beads that became smaller and smaller over time. The loss of water from the confined environment in which the enzyme is located led to the observed slowdown of the reaction. This hypothesis was confirmed by experiments performed by varying the amount of water inside the reaction medium; the time required for complete hydrolysis of the p-NPA was decreased from 24 to 4 h with a water percentage of 5% and 75% respectively (Fig. S6, ESI†).
Therefore, given all the considerations made so far, the Beads 1 formulation, being one of the fastest preparations and ensuring better mass transfer, thus allowing their possible use with a wide range of substrates, has been chosen to perform thermostability tests and kinetic resolution of (R,S)-1-phenylethyl acetate.
![]() | ||
Fig. 6 Thermal stability of free (![]() ![]() ![]() ![]() |
The effect of immobilization on Ca-alginate beads on CRL stability is clearly highlighted by the figure at both investigated temperatures. Immobilized CRL activity at 25 °C did not decrease after 7 days of incubation while the free form lost 30% of its initial activity after 5 days and 60% after one week. Even more evident was the stabilization effect at 50 °C. In fact, the residual activity of free lipase, after only 8 h, was lower than 40% and continued to decrease over time until reaching a value of about 20% after one week of incubation. On the other hand, the remaining activity of immobilized CRL was about 70% after 1 day of heat treatment and remained unchanged throughout the week.
Moreover, stability tests were carried out with Beads 6 at 50 °C to assess the effect of their higher compactness on the enzyme thermal stability. They showed that the different operational conditions used in the preparation of the two formulations did not affect the stability of the enzyme, being the loss of activity after 24 h equal to 66% (data not shown).
The data reported in Fig. 6 show the trends of residual activity over time, but do not consider the differences in the reaction rate before incubation. In particular, hydrolysis rate with free CRL was 30% higher than that of immobilized one both at 25 °C (183 vs. 132 μM min−1) and 50 °C (520 vs. 370 μM min−1). This effect could be due to a slower diffusion of the substrate inside the support. However, after a week of incubation the reaction rate of Beads 1 at 25 °C and 50 °C was 2 times higher than that of the free enzyme, thanks to the improvement in enzymatic stabilization following the entrapment.
Given the excellent recyclability and thermal stability of Beads 1, and therefore their possible application in industrial processes, they were used in preliminary tests for the resolution of (R)-1-phenylethanol in aqueous solution starting from racemic 1-phenylethyl acetate as model substrate. In the literature, several papers report about the stereoselective kinetic resolution of rac-1-phenylethyl acetate catalyzed by other lipases,69–73 in which very high enantiomeric excess towards the (R)-enantiomer were obtained, but with not too satisfying yields. More recently, marine microbial GDSL lipase MT6 showed opposite stereoselectivity, as it hydrolyzed racemic 1-phenylethyl acetate to generate (S)-1-phenylethanol instead of (R)-1-phenylethanol.74 During reaction course, as the conversion increased, the ee of the product (eep) decreased and, at the optimal reaction time (12 h), a conversion of 28.5% with an ee value of the substrate (ees) higher than 97% was obtained. On the other hand, CRL solubilized in phosphate buffer at pH 7.2 showed no enantiopreference for the (R)-acetate with an ee value of only 44% and an enantioselectivity factor of 4.75
Here, the course and selectivity of the kinetic resolution of (R,S)-1-phenylethyl acetate were checked, as reported in materials and methods, by chiral HPLC and all the chromatograms are shown in Fig. S7 (ESI†). The percentage of enantiomeric excess (ees and eep), conversion (c) as well as enantioselectivity (E), determined using eqn (3)–(6), are reported in Table 4.
Reaction time (h) | ee s, % | ee p, % | c, % | E |
---|---|---|---|---|
8 | 3.5 | 61.5 | 5.3 | 4.3 |
24 | 17.6 | 62.8 | 21.9 | 5.2 |
32 | 18.1 | 61.9 | 22.6 | 5.1 |
48 | 23.1 | 64.5 | 26.4 | 5.8 |
72 | 49.5 | 61.7 | 44.5 | 6.8 |
96 | 83.3 | 60.9 | 57.8 | 10.3 |
As observed from these results, while ees increased with reaction time and conversion, eep remained almost constant. After 96 h of reaction, the highest values of enantioselectivity (E = 10.3), conversion (c = 57.8%) and enantiomeric excess of substrate (ees = 83.3%) were observed. These values are slightly higher but do not differ so much from those reported in the literature regarding free lipase in phosphate buffer, which showed an ees value of 44% and an enantioselectivity factor of 4.75 Given the intrinsic chirality of the matrix, we performed these tests to assess whether the immobilization improved the enantioselectivity of the enzyme towards the chosen substrate. This has not happened, but the entrapment of CRL improved its stability under non-physiological conditions and temperatures and made it recyclable for multiple cycles, which is essential for industrial applications.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2nj01160c |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2022 |