You are on page 1of 13

Chemical Engineering Journal 459 (2023) 141516

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Polydopamine-assisted one-step immobilization of lipase on α-alumina


membrane for fouling control in the treatment of oily wastewater
Jéssica Mulinari a, b, c, *, Alan Ambrosi b, Yuren Feng c, d, Ze He c, d, Xiaochuan Huang c, d,
Qilin Li c, d, Marco Di Luccio b, Dachamir Hotza a, J. Vladimir Oliveira a
a
Graduate Program in Chemical Engineering (PósENQ), Department of Chemical Engineering and Food Engineering (EQA), Federal University of Santa Catarina (UFSC),
Trindade, Florianópolis, Santa Catarina 88040-900, Brazil
b
Laboratory of Membrane Processes (LABSEM), Graduate Program in Food Engineering, Department of Chemical Engineering and Food Engineering (EQA), Federal
University of Santa Catarina (UFSC), Trindade, Florianópolis, Santa Catarina 88040-900, Brazil
c
Department of Civil and Environmental Engineering, Rice University, MS 519, 6100 Main Street, Houston, TX 77005, USA
d
Nanotechnology-Enabled Water Treatment Center (NEWT), Rice University, MS 6398, 6100 Main Street, Houston 77005, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Covalent enzyme immobilization is generally a time-consuming and multistep procedure that uses toxic solvents
Eversa transform 2.0 and requires more than one chemical, making industrial upscaling unattractive. Using an aqueous polydopamine
Enzyme immobilization (PDA) solution for enzyme immobilization is a greener alternative. Usually, enzyme immobilization using PDA is
Ceramic membrane
performed in two steps: dopamine polymerization on the material surface followed by enzyme immobilization. A
Reactive membrane
Polydopamine
few recent studies applied a one-step strategy by mixing dopamine and enzyme in the coating solution, reducing
One-pot immobilization the immobilization time, chemical consumption, and wastewater generation. This study compares the two-step
and one-step approaches to immobilizing the lipase Eversa Transform 2.0 (ET2) on an α-alumina membrane. The
one-step immobilization method achieved similar enzyme loading, membrane hydrolytic activity, and enzyme-
specific activity to those of the two-step method. The ET2 immobilized using both strategies showed excellent
fouling resistance and self-cleaning performance in oily wastewater filtration. The membrane modified by the
one-step approach exhibited a lower reduction in pure water permeance after oil fouling (35%) and a higher
permeance recovery (90%) than the one modified by the two-step method (40% and 74%, respectively). This
better performance can be due to the higher hydrophilicity of the modified membrane and higher stability over
reaction time shown by the enzyme immobilized by the one-step strategy. The higher stability can be attributed
to more attachment points between the enzyme and PDA, increasing the enzyme rigidity and preventing
conformational changes.

1. Introduction Enzymes have been immobilized on different materials. Polymeric


materials are most often used as they possess reactive surface functional
Enzymes are used in a variety of applications as catalysts [1,2]. groups that can bond to the enzymes [6,7]. However, inorganic mate­
Because enzymes are water-soluble macromolecules, additional sepa­ rials have important advantages as a support, including better me­
ration and purification processes are needed to recover and reuse the chanical, chemical, and thermal stability [8,9]. Among inorganic
enzymes, which increase the complexity of the system and raises the materials, ceramics stand out for their relatively low cost, long service
cost. One strategy is to immobilize enzymes on solid supports, allowing life, and easy pore size control during fabrication [8,10].
their reuse in multiple reaction cycles. Furthermore, enzyme immobi­ Enzymes can be immobilized by physical methods, such as adsorp­
lization enables continuous operation and, in some cases, increases tion, entrapment, and encapsulation, or chemical methods, such as co­
enzymatic stability and activity, improving the process’s economics valent bonding. When immobilized by adsorption, the enzyme forms
[3–5]. weak interactions with the support (hydrogen bonds, van der Waals

* Corresponding author at: Department of Chemical and Food Engineering, Federal University of Santa Catarina, Caixa Postal 476, Florianópolis, Santa Catarina
88040-900, Brazil.
E-mail address: jessica.mulinari@posgrad.ufsc.br (J. Mulinari).

https://doi.org/10.1016/j.cej.2023.141516
Received 3 November 2022; Received in revised form 22 December 2022; Accepted 17 January 2023
Available online 20 January 2023
1385-8947/© 2023 Elsevier B.V. All rights reserved.
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

forces, hydrophobic interactions, ionic bonding, or affinity binding) and Only three studies were found in the literature about lipase immo­
so can easily desorb due to changes in pH, ionic strength, temperature, bilization on membranes to develop oil fouling-controlling and self-
pressure, etc. [2,11]. Entrapment and encapsulation of the enzyme cleaning abilities: Schulze et al. [24], Schmidt et al. [23], and Schmidt
within the support can cause massive mass transfer limitations, pre­ et al. [26]. The results presented by these studies are promising, with the
venting contact between the substrate and the active site [11,12]. Co­ biocatalytic membrane showing permeance reductions after emulsion
valent bonding is applied when strong interactions between the enzyme filtration as low as 40 % (the pristine membrane had a 60 % reduction)
and the support are necessary to prevent leakage, generally resulting in and permeance recoveries as high as 100 % after cleaning just using a
higher enzyme stability and reusability compared to other methods buffer at proper pH and temperature. However, all three studies used
[2,13,14]. polymeric flat-sheet membranes. So far, most of the works using
However, to have covalent bonding between the support and the immobilized enzymes to decrease membrane fouling focused on
enzyme, both have to contain functional groups. Since ceramics lack biofouling control [27–29] or protein fouling mitigation [25,30,31].
reactive surface functional groups, they have to be functionalized before Moreover, flat-sheet polymeric membranes are the most evaluated
the immobilization. Silicon-based polymers, such as organosilanes, are [23,24,26], and food industry processes are the most studied [30,32,33].
the most used agents to functionalize the ceramic. The organosilane Thus, it is noted that the use of enzymes as antifouling and self-cleaning
reacts with the hydroxyl functional groups on the ceramic surface, agents in membranes is a relatively unexplored area due to the huge
generating organic functional groups available for further reactions with variety of enzymes, membranes, and membrane separation processes
the enzyme [8]. The most common chemicals used to covalently that the use of immobilized enzymes can improve.
immobilize enzymes on ceramic materials are 3-aminopropyltriethoxy­ Compared to other membrane modification methods to decrease
silane (APTES) as a functionalization agent and glutaraldehyde as an fouling, using immobilized enzymes is the simplest way to provide the
activating agent [8,15]. Nevertheless, the immobilization procedures, membrane with a self-cleaning capacity as well. Photocatalytic mem­
generally requiring various chemicals and multiple steps, are time- branes can also degrade the fouling during the cleaning; however, the
consuming and involve toxic solvents, making industrial upscaling performance is limited on how much light reaches the membrane sur­
unattractive. face. Usually, a UV light source is required since photocatalysis is limited
Therefore, simpler immobilization protocols using fewer chemicals under visible light, which increases the energy demand [34]. Electro­
are essential to make ceramic materials more competitive as enzyme catalysis is also an alternative in which the membrane acts as an anode
supports. Polydopamine (PDA) can strongly adhere to different mate­ or cathode; however, a current source is necessary, also increasing the
rials [16,17] and has therefore been used for enzyme immobilization. energy demand [35]. Fenton-like processes by immobilizing iron on the
Due to the presence of a large number of reactive functional groups (e.g., membrane are also possible, but they require hydrogen peroxide to
amine, imine, quinone, and catechol), PDA coating can bond enzymes to achieve improved performances [36]. Moreover, photocatalysis, elec­
ceramic materials [18–22]. For example, the amino groups located at trocatalysis, and Fenton-like processes are not specific and can promote
the side chain amino acids of the enzyme can react with the quinone several side reactions, which is not a problem when enzymes are used
groups of the PDA coating through the Michael addition or Schiff base due to their high specificity [37].
reaction, covalently bonding the enzyme to the ceramic surface. Similar Therefore, besides developing a one-step method to immobilize the
to the conventional method, enzyme immobilization using PDA is usu­ lipase on the membrane, this study also aimed to develop a catalytic
ally performed in two steps: dopamine polymerization on the material membrane with fouling control and self-cleaning properties to be
surface in an alkaline environment, followed by enzyme immobilization applied in the filtration of oily wastewater. If successful, the lipase-
using standard buffers. Our previous study compared the PDA coating active membranes can be applied in the wastewater treatment of
method on ceramic membranes with traditional enzyme immobilization various food industries (from the production of poultry, beef, and fish, to
techniques such as silanization [21]. The PDA coating method resulted dairy, cheese, edible oils, and prepared food).
in a higher enzyme loading and membrane hydrolytic activity than the
traditional salinization method [21]. 2. Materials and methods
Although the use of nature-inspired PDA represents an
environmentally-friendly and simpler alternative to conventional 2.1. Materials and chemicals
immobilization agents, it is still a time-consuming multistep method that
generates a large volume of wastewater. Here, we adopted a one-step Purified and concentrated lipase Eversa Transform 2.0 (ET2) pro­
coating strategy to reduce the immobilization time, chemical usage, duced by a genetically modified strain of Aspergillus oryzae was used
water consumption, and wastewater generation. Therefore, if the one- (Novozymes, Denmark). The commercial enzymatic solution was dia­
step method using PDA proves to be a good alternative to the two-step lyzed for 120 h using a cellulose membrane (12–14 kDa) and sodium
approach, besides being a greener technique and easier to scale up, phosphate buffer at 50 mmol⋅L− 1 and pH 6.0. The buffer was replaced
the costs of the immobilization process can be lower, making it more every 12 h. The purified enzyme solution was frozen at − 50 ◦ C and
industrially attractive than the traditional immobilization protocols. lyophilized for 48 h (Liotop model L101n, Brazil). The resulting powder
Membrane industries can benefit from this technology as well as was stored at 4 ◦ C until use.
research institutions, since, due to the non-specificity of the PDA Tubular α-alumina membranes were custom-made (Tecnicer, Brazil).
coating, this simpler immobilization technique can be used to modify a They are 25 cm in length, 1.2 cm in outer diameter, and 0.8 cm in inner
wide range of materials using different enzymes. diameter. The total porosity of the membranes is 30 ± 2 % and it was
In the case of oily wastewater treatment, the enzyme lipase is a good calculated using the apparent density (2.8 ± 0.1 g⋅cm− 3) determined by
candidate to be immobilized on the membrane since they can break the Archimedes’ principle using deionized water [38] and the absolute
triacylglycerols of the oil into free fatty acids and glycerol. Therefore, density (4.0 ± 0.1 g⋅cm− 3) determined by helium pycnometry (Quan­
when the biocatalytic membrane is used to filtrate oily wastewater, the tachrome Ultra pycnometer 1000, EUA). The average pore size deter­
immobilized lipases can degrade the oil fouling during the filtration and mined by mercury intrusion porosimetry (Micromeritics Poresizer
cleaning procedure. The hydrolysis of the fouling layer during the model 9320, USA) was 290 nm [21].
cleaning can be performed through a simple activation by adjusting the Analytical grade dopamine hydrochloride (DA, 98 %, C8H11NO2.
reaction conditions, such as pH and temperature [23–25]. Thus, the HCl, Sigma-Aldrich, USA), sodium dodecyl sulfate (SDS, ≥90 %,
immobilized enzymes provide the membrane with a self-cleaning ca­ C12H25SO4Na, Synth, Brazil), polyvinyl alcohol (PVA, 98 %, (C2H4O)n,
pacity, decreasing or even avoiding chemical cleaning, which makes the Mw = 104.5 kg/mol, NEON, Brazil), sodium hydroxide (≥97 %, NaOH,
process greener and more sustainable. NEON, Brazil), ethanol (99.5 %, C2H6O, NEON, Brazil), and

2
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

phenolphthalein (99 %, C20H14O4, Sigma-Aldrich, USA) were used as adding ethanol right before the addition of the enzyme. The specific
purchased. Tris(hydroxymethyl)aminomethane (≥99.5 %, C4H11NO3, hydrolytic activity (HAS) was calculated using Eq. (1). All tests were
NEON, Brazil) and hydrochloric acid (37 %, HCl, NEON, Brazil) were performed in triplicate.
used to prepare the 50 mM Tris-HCl buffer (pH 8.5) and 100 mM Tris-
(Vs − Vc ) × M × 1000
HCl buffers (pH 9 and 10). The 100 mM sodium phosphate buffers HAS (μmol/min.g) = (1)
t × V × PC
(pH 6, 7, and 8) were prepared using sodium phosphate monobasic (98
%, NaH2PO4, NEON, Brazil) and sodium phosphate dibasic (P.A., where Vs is the NaOH volume used to titrate the sample (mL), Vc is the
Na2HPO4, NEON, Brazil). The 100 mM potassium acetate buffers (pH 4 NaOH volume used to titrate the control assay (mL), M is NaOH molarity
and 5) were prepared using glacial acetic acid (99.8 %, CH3CO2H, (mol⋅L− 1), t is the reaction time (min), V is the volume of enzymatic
NEON, Brazil) and potassium acetate (≥98.5 %, CH3CO2K, NEON, solution added to the reaction media (mL), and PC is the protein content
Brazil). Refined soybean oil (Soya®) was purchased from a local market. of the enzymatic solution (g⋅mL− 1) determined by the Bradford method
using bovine serum albumin as a standard [40].
2.2. ET2 hydrolytic activity
2.3. ET2 immobilization
The ET2 hydrolytic activity was determined using soybean oil as the
substrate. Soybean oil emulsion (25 % v⋅v− 1) was prepared immediately Before each test, the ends of each membrane were sealed with a
before each experiment by emulsifying soybean oil in a 2 % (m⋅m− 1) thermoplastic adhesive to prevent the solution from entering the inside
aqueous solution of PVA using a magnetic stirrer at 1500 rpm for 10 min of the tubular membrane. PDA deposition was performed by air oxida­
according to the method proposed by Fu et al. [39]. Then, 4 mL of the tion [21,41,42]. For the two-step approach (Fig. 1a), the membrane was
soybean emulsion was added to 6 mL of sodium phosphate buffer (100 immersed in a solution of dopamine hydrochloride (DA) (2 mg⋅mL− 1) in
mmol⋅L− 1, pH 7) with 1 mg⋅mL− 1 of the lyophilized ET2 (volume ratio Tris-HCl buffer (pH 8.5, 50 mmol⋅L− 1) for 6 h at room temperature with
soybean oil emulsion:buffer of 2:3, resulting in a soybean oil concen­ gentle stirring at 40 rpm. Then, the membrane was rinsed with deionized
tration of 90 g⋅L− 1). The buffer solution with ET2 was stirred at 300 rpm water and immersed in a 2 mg⋅mL− 1 ET2 solution in Tris-HCl buffer (pH
for 1 h before being mixed with the soybean oil emulsion. After incu­ 8.5, 50 mmol⋅L− 1) at 40 rpm for 12 h. Before the immobilization, the
bation for 30 min at 40 ◦ C and agitation at 160 rpm, the reaction was ET2 solution was stirred for 1 h at 300 rpm to break possible enzyme
interrupted by adding 15 mL of ethanol. The amount of fatty acids agglomerates. For the one-step strategy (Fig. 1b), a solution of 2
liberated during the reaction was determined by titration using NaOH mg⋅mL− 1 of ET2 in Tris-HCl buffer (50 mmol⋅L− 1 pH 8.5) was stirred for
0.05 mol⋅L− 1 and phenolphthalein. Control assays were carried out by 1 h, and then 2 mg⋅mL− 1 of DA was added to the solution. The

Fig. 1. Schematic representation of the immobilization methods tested in this study: (a) two-step immobilization, (b) one-step immobilization, and (c) adsorption.

3
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

membrane was immersed in the DA/ET2 solution for 12 h at 40 rpm. Table 1


Adsorption of ET2 in a pristine membrane (Fig. 1c) was also tested by Comparative of the two-step and the one-step methods for ET2 immobilization
immersing the membrane in a 2 mg⋅mL− 1 ET2 solution in Tris-HCl buffer on the ceramic membrane using PDA (approximate values based on the modi­
for 12 h. After all the immobilization procedures, the membranes were fication methods used in this study).
washed with sodium phosphate buffer (100 mmol⋅L− 1 pH 7) to remove Method Modification Chemicals usage Wastewater
loosely bound enzymes until no protein was detected using the Bradford time (h) generation (L⋅m− 2)
DA (g⋅ Tris-HCl 100
method. Triplicate samples were prepared using each immobilization m− 2) mM pH 8.5 (L⋅
method. m− 2)
The amount of enzyme immobilized on the membrane was deter­ Two- 18 30 30 30
mined by the change in ET2 concentration in the solution before and step
after the immobilization, as measured by the Bradford method, also One- 12 30 15 15
step
accounting for the enzyme removed during the rinsing step. Enzyme
loading (EL) was calculated according to Equation (2).
( )
( )
Co − Cf × V − Mr immobilized ET2 were investigated during oil hydrolysis. Both free and
EL g⋅m− 2 = (2) immobilized ET2 activities were determined at different temperatures
A
(20 – 50 ◦ C) and pH (4 – 10). The hydrolysis activity was then evaluated
where C0 is the initial enzyme concentration in the solution (before the for 24 h using the optimal temperature and pH. Free and immobilized
immobilization) (g⋅mL− 1), Cf is the final protein content (after the ET2 activities were determined as described in Sections 2.2 and 2.3,
immobilization) (g⋅mL− 1), V is the volume of enzyme solution used in respectively; however, due to the different pH values used, instead of
the test (mL), Mr is the amount of loosely bound enzyme removed by the phenolphthalein as an indicator, titration was performed until pH 11,
rinsing step (g), and A is the membrane projected external area (m2). and control assays were performed for each pH [43]. The effect of pH,
The hydrolytic activity of the immobilized ET2 was evaluated by the temperature, and reaction time were evaluated according to a relative
same method used to determine the free ET2 activity. For this, a test activity (%), calculated by normalizing the obtained activities by the
solution was prepared using the same soybean oil emulsion:buffer vol­ highest one.
ume ratio of 2:3 and final soybean oil concentration of 90 g⋅L− 1 (spe­
cifically, 50 mL of soybean emulsion was mixed with 75 mL of sodium 2.5. Membrane characterization
phosphate buffer (100 mmol⋅L− 1 pH 7)). The membranes were
immersed in the test solution and incubated at 40 ◦ C for 30 min with The surface morphology and elemental composition of the pristine
gentle mixing at 40 rpm. Then, a homogenized sample of 10 mL was and modified membranes were characterized by Scanning Electron Mi­
withdrawn from the solution and titrated with 0.05 mol⋅L− 1 NaOH to croscopy (SEM, FEI Helios NanoLab 660 Dual, USA) coupled with
determine the amount of free fatty acids formed from the hydrolysis Electron Dispersion X-ray (EDX). Membrane hydrophilicity/hydropho­
reactions. Control assays were performed under the same conditions bicity was determined using water and n-heptane vapor adsorption
using membranes subjected to the same immobilization protocols but measurements as described by Mulinari et al. [21] and adapted from
without the ET2 enzyme. The membrane hydrolytic activity (MHA) and Nishihora et al. [44] and Prenzel et al. [45]. Membrane surface chem­
the specific hydrolytic activity of the immobilized ET2 (MHAS) were istry was characterized by X-ray Photoelectron Spectroscopy (XPS, PHI
determined according to Equations (3) and (4), respectively. Quantera, MN, USA) using monochromatic Al Kα X-rays. Since alumina
( ) (Vs − Vc ) × M × (Vt /V) × 1000 is an insulating material, the spectra of all samples were charged
MHA μmol⋅min− 1 .m− 2
= (3) correctly by shifting all peaks to the adventitious C 1 s spectral
t×A
component (C–C, C–H) at 284.8 eV. The membranes and the lyophi­
( ) (Vs − Vc ) × M × (Vt /V) × 1000 lized ET2 were also analyzed by attenuated total reflectance Fourier
MHAS μmol⋅min− 1 .g− 1 = (4)
t × A × EL transform infrared spectroscopy (ATR-FTIR, Cary 660, Agilent Tech­
nologies, CA, USA).
where Vs is the NaOH volume used to titrate the sample (mL), Vc is the
NaOH volume used to titrate the control assay (mL), M is NaOH molarity
(mol⋅L− 1), Vt is the total volume of emulsion used in the test (mL), V is 2.6. Oil emulsion filtration tests
the volume of the emulsion sample titrated (mL), t is the reaction time
(min), A is the membrane projected external area (m2), and EL is the The ET2-immobilized membranes were tested for their fouling
enzyme loading on the membrane (g⋅m− 2). resistance and self-cleaning function in the filtration of a soybean oil
Residual activity, as defined by Eq. (5), was used to estimate the emulsion following previously reported protocols [21,23,41]. Since the
effect of immobilization on enzyme activity. immobilization was carried out on the membrane’s outer surface, the
filtration was performed outside-in using the filtration system schema­
Residual activity (%) = 100 ×
MHAS
(5) tized in Fig. 2. To prepare the test solution, soybean oil (1 g⋅L− 1) was
HAS emulsified in a 2 mmol⋅L− 1 SDS aqueous solution by magnetic stirring at
Analysis of variance (ANOVA) and Tukey’s test (p < 0.05) were 1000 rpm for 5 min followed by probe sonication (Unique model
performed on enzyme loading, membrane hydrolytic activity, and spe­ DES5000, Brazil) at 100 W for 5 min. In the filtration experiment,
cific hydrolytic activity. Table 1 compares the two-step and one-step deionized water was first filtered at 1 bar and 1 L⋅min− 1 of retentate flow
immobilization methods regarding immobilization time, consumption rate to determine the initial pure water permeance of the membrane.
of chemicals, and wastewater generation per membrane area. If suc­ Then, two filtration steps of 2 h each were performed using the soybean
cessful, the one-step strategy can be less time-consuming (–33 %) and oil emulsion. At the end of each filtration step, a backwash was carried
have a lower generation of wastewater (-50 %) and consumption of out for 5 min at 1 bar and 1 L⋅min− 1 of retentate flow rate using
chemicals (-25 %). deionized water. After the final backwash, pure water permeance was
measured again to evaluate the degree of fouling. All the steps of the
experiment were performed in crossflow mode with complete recircu­
2.4. Operational stability of the immobilized and free ET2 lation of the retentate back to the feed reservoir. The transmembrane
pressure used was 1 bar, the retentate flow rate was 1 L⋅min− 1, and the
The effects of pH, temperature and reaction time on the free and crossflow velocity was 0.20 m⋅s− 1. Sealing rings were fixed at 1 cm from

4
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Ren et al. [22] when comparing adsorption with a two-step PDA method
for immobilization of Candida rugosa lipase type VII on iron oxide
nanoparticles: the use of PDA tripled the enzyme loading. However, the
formation of covalent bonds between ET2 and PDA decreased the spe­
cific activity of the immobilized ET2 compared to the adsorption
method, as evident in the lower residual enzyme activity resulting from
the two PDA immobilization methods. This is attributed to structural
changes or denaturing of the enzyme as well as blockage of active sites
by PDA. Adsorption, on the other hand, involves weaker, non-specific
interactions, causing fewer or no changes in the enzyme structure
[2,12]. Nevertheless, all immobilization methods caused a large drop in
enzyme activity compared to the free enzyme, as indicated by the low
residual activity.
When using the one-step strategy, there was a concern that the PDA
might cover the enzyme and make the active site less accessible to the
reactant. However, our results show that the two methods resulted in
similar protein loading and hydrolytic activity; the Tukey test did not
show a statistically significant difference between the two approaches.

Fig. 2. Schematic representation of the filtration system used for the oil
emulsion filtration tests. 3.2. Free and immobilized ET2 optimum operating pH and temperature

each end of the membrane, resulting in an effective membrane length of The stability of enzymes as a function of temperature, pH, and
23 cm with an effective area of 79.5 cm2. After the filtration experi­ operating time is an important factor in determining their suitability for
ments, the membranes were immersed in sodium phosphate buffer (100 applications that may encounter fluctuation of temperature and solution
mmol⋅L− 1 pH 7) for 12 h at 40 ◦ C with stirring at 40 rpm to evaluate the pH, and require a long lifetime of the enzyme. Fig. 3 shows the activity of
membrane’s self-cleaning capacity. Finally, the membranes were rinsed free and immobilized ET2 as a function of temperature, pH, and time.
with deionized water, and the water permeance was measured again. Regardless of the method used, immobilization did not alter the en­
The pristine membrane and membranes functionalized with and without zyme’s optimal operating temperature (around 40 ◦ C), but it increased
the enzyme were tested in triplicate. Oil concentration in the feed and ET2 activity at other temperatures, especially at temperatures higher
permeate was monitored throughout the experiment using a UV–visible than the optimal temperature (Fig. 3a). In the temperature range of 40 to
spectrophotometer (Shimadzu model UV-2550, Japan) at a wavelength 60 ◦ C, the one-step immobilization method yielded significantly higher
of 240 nm. enzyme activity than the two-step approach. This improved thermo­
Three cycles of filtration and cleaning were performed using the stability can be attributed to a larger number of attachment points be­
same membrane to investigate their reusability in long-term applica­ tween the enzyme and PDA (as schematized on Fig. 7b). As the enzyme
tions. The number of filtration steps within each cycle was two. The moieties involved in the immobilization reaction are fixed in location,
emulsion was prepared as described above as well as the filtration and they increase the enzyme rigidity and hinder conformational changes
self-cleaning experiments. when temperature increases [4,20].
Fig. 3b shows that ET2 immobilization using PDA coating increased
3. Results and discussion the enzyme’s optimum operating pH from pH 6 for the free enzyme to
pH 7 and 9 for the one-step and two-step PDA-immobilized enzymes,
3.1. ET2 immobilization respectively. The immobilization resulted in lower enzyme activity at
pH <7 as well as pH 10 but increased its activity in the pH range of 7 to
The lyophilized ET2 was immobilized by PDA coating using the two- 9. Touqeer et al. [46] observed a similar change after immobilizing
step and one-step methods and by simple adsorption in the membrane Aspergillus terreus lipase on PDA-coated iron oxide nanoparticles: the
for comparison. Table 2 shows the results of enzyme loading, membrane maximum enzyme activities were obtained at more alkaline pH values
hydrolytic activity, specific hydrolytic activity, and residual activity. after immobilization.
The residual activities were calculated based on the specific activity of The lipase immobilized by the one-step method showed higher ac­
free lyophilized ET2: 1391.2 µmol⋅min− 1⋅g− 1. tivities than the one immobilized by the two-step method for all the pH
Immobilization by simple adsorption (using the membrane with no values tested except pH 9. It is hard to determine exactly why this
functionalization) resulted in a very low enzyme loading on the mem­ happened. It is an agreement in the literature that immobilization can
brane (0.9 ± 0.2 g⋅m− 2) and, consequently, a low membrane hydrolytic help to improve stability in the microenvironment near the immobilized
activity (686 ± 80 µmol⋅min− 1⋅m− 2). The use of PDA greatly increased enzyme [47]. So, in this case, although both methods used PDA as a
the enzyme loading on the membrane for both the two-step (3.1 ± 0.2 bonding agent, they resulted in slightly different coatings, as demon­
g⋅m− 2) and the one-step (3.1 ± 0.1 g⋅m− 2) approaches, which resulted in strated by the XPS analysis (Fig. 6), which can indicate some variances
higher membrane hydrolytic activities. Similar results were obtained by in the microenvironment near the enzyme and, therefore, different be­
haviors at the same pH values.

Table 2
Enzyme loadings (EL) and hydrolytic activities of the ET2 immobilized on the α-alumina membranes by different protocols.
Immobilization strategy EL (g⋅m− 2) MHA (µmol⋅min− 1⋅m− 2) MHAS (µmol⋅min− 1⋅g− 1) Residual activity (%)
a a a
Adsorption 0.91 ± 0.16 685.6 ± 80.0 755.1 ± 47.4 54.3 ± 3.4a
Two-step PDA 3.12 ± 0.23b 1822.8 ± 77.7b 587.8 ± 59.6b 42.3 ± 4.3b
One-step PDA 3.10 ± 0.10b 1986.2 ± 40.0b 641.3 ± 17.8b 46.1 ± 1.3b

All values are averages of triplicate tests and their respective standard deviations. Values in the same column with different superscripts indicate statistically significant
differences according to Tukey’s test (p < 0.05).

5
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Fig. 3. Relative activity as a function of (a) temperature, (b) pH, and (c) operational time of the free and immobilized ET2 by the two-step and one-step approaches
using PDA.

Fig. 4. SEM images of the active surface of the (a) pristine membrane, (b) PDA-coated membrane, (c) membrane with ET2 immobilized by the two-step method, and
(d) membrane with ET2 immobilized by the one-step method; (e) elemental analysis of the membranes by EDX.

Fig. 3c shows the enzyme stability over time for the free and highest stability, which is consistent with the larger number of attach­
immobilized ET2. Both immobilization methods increased the opera­ ment points and, therefore, higher rigidity of the enzyme.
tional stability of the enzyme, showing superior relative activity over 4 h
of the hydrolysis reaction. The free and immobilized ET2 had similar 3.3. Membrane characterization
relative activities in the first two hours, consistent with the MHA results
reported in Table 2 for 30 min of reaction. After 2 h, the relative activity Fig. 4 shows the SEM images of the pristine membrane as well as
of free ET2 dropped much faster than those immobilized by either membranes after different coating steps. Analyzing the SEM images, the
method. The enzyme immobilized by the one-step strategy showed the membrane modified by the one-step strategy (Fig. 4d) showed a more

6
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Fig. 5. ATR-FTIR of the pristine membrane, PDA-coated membrane, mem­


branes with immobilized ET2 by the two-step and one-step methods, and
lyophilized ET2.

uniform coating than the one modified by the two-step method (Fig. 4c),
which can be a result of different PDA polymerization times (6 h of
polymerization for the two-step modified membrane and 12 h of
simultaneous polymerization and immobilization for the one-step
modified membrane). EDX analysis showed an increase in the carbon
content of the membrane surface after enzyme immobilization: from 20
% for the PDA-coated membrane to 24 % and 28 % for the membranes
with ET2 immobilized by the two-step and one-step method, respec­
tively. The EDX mapping for each membrane (Fig. S1) shows a homo­
geneous distribution of carbon and oxygen on the membrane surface,
suggesting that the coating was uniform. Fig. S2 shows the photograph
of the pristine and the one-step modified membrane. As can be seen,
polydopamine turns brown after polymerization in an alkaline solution.
ATR-FTIR analyses (Fig. 5) show no significant difference in surface
chemistry of the ET2 immobilized membranes obtained using the two-
step and one-step methods, which corroborates the similar hydrolytic
activities obtained using these membranes. The membranes with
immobilized ET2 also presented similar functional groups found in the
free lipase sample: O–H stretching (3280 cm− 1), a combination of N–H
stretching from amine groups, and C–H stretching from alkenes (2920
cm− 1), C–– O stretching of primary amides from the α-helix secondary
assignment, common in the secondary structure of proteins (1650
cm− 1), and N–H bending of secondary amides (1540 cm− 1) [48,49]. It
also showed a signal at 1050 cm− 1, corresponding to C–N stretching
from amines, present at both enzyme and PDA coating [50,51].
Fig. 6a and Table 3 show the surface chemical composition of the
membranes and the free ET2 characterized by XPS analysis. Almost no
aluminum was detected after the membrane modification while the
carbon content increased, indicating a successful modification of the
membrane surface. Moreover, nitrogen peaks appeared after PDA
coating and increased after ET2 immobilization. Table 3 shows the
relative amounts of the elements detected by the XPS analysis. After ET2 Fig. 6. (a) XPS spectra of the pristine membrane, modified membranes, and
immobilization, the N/C atomic ratio increased from 0.12 ± 0.01 (PDA- free ET2, (b) fitted carbon (C 1 s), and (c) nitrogen (N 1 s) spectra for the
coated membrane) to 0.20 ± 0.01 for the two-step method and 0.19 ± membranes with PDA coating and with immobilized ET2 by the two-step and
one-step methods.
0.03 for the one-step approach, which corroborates the similar results
for enzyme loading in both membranes.
Fig. 6b shows the fitted carbon spectra of the modified membranes. the alkali environment (pH 8.5) as describe by Lu and Yu [52] and the
The ET2-immobilized membranes exhibit an increase in the content of condensation reaction of the carboxyl groups in the enzyme with the
C–– O or C–– N and a decrease in C–H and C–C contents, suggesting a amino groups of the PDA (Fig. 7a). Wang et al. [53] also noticed an
Schiff base reaction between the quinone groups of PDA and the amino increase in C–– O after immobilizing perhydrolase on the PDA coating
groups of ET2, as schematized on Fig. 7a. It may also be attributed to the and attributed it to the formation of protein-PDA complexes.
oxidation of the remaining PDA catechol groups to quinone groups in The N 1 s spectrum (Fig. 6c) shows a considerable increase in the
intensity of R2NH binding energy when ET2 was immobilized on the

7
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Fig. 7. (a) Proposed immobilization mechanism for both two-step and one-step methods: Schiff base reaction, Michael addition and condensation reaction; (b)
multipoint attachment occurring during the one-step modification and partial covering of PDA on the enzyme surface. Adapted from Mulinari et al. (2022).

Fig. 8 compares the hydrophilicity/hydrophobicity of the mem­


Table 3
branes. As expected, water vapor adsorption by all the membranes was
Elemental composition of free ET2 and pristine, PDA modified, and ET2-
much greater than n-heptane vapor adsorption due to the hydrophilic
immobilized membranes analyzed by XPS.
nature of alumina. Consistent with our previous observation [21], PDA
Sample Atomic composition (%)
coating reduced both water and n-heptane vapor adsorption, which can
C N O Al be a result of pore blockage. Due to the hydrophilic character of PDA, the
Pristine 17.78 ± – 62.83 ± 19.39 ± decline of n-heptane vapor uptake by the PDA-coated membrane was
3.27 3.24 0.03 more intense than that of water, resulting in a higher water/n-heptane
PDA 64.05 ± 7.56 ± 0.13 26.41 ± 1.99 ± 1.13 ratio. On the contrary, the addition of ET2 to the membrane surface
3.60 2.60
greatly increased the adsorption of both vapors, which can be attributed
PDA + ET2 (two- 60.45 ± 11.86 ± 25.12 ± 0.96 ± 0.29
step) 0.46 0.41 0.48 to the enzyme’s amphiphilic nature [54]. The results also show that the
PDA + ET2 (one- 58.71 ± 12.68 ± 27.15 ± – water vapor adsorption is consistent with the enzyme loading on the
step) 1.54 1.94 0.25 membrane since the membranes with ET2 immobilized through PDA
ET2 57.10 ± 6.80 ± 0.73 32.25 ± – coating showed higher values. The slightly higher water vapor uptake
0.91 1.84
presented by the membrane modified by the one-step method can be a
consequence of a higher quantity of polar covalent bonds on the mem­
membrane, suggesting that the amino groups of the enzyme could have brane surface, such as C-OH, C–N, and R2-NH, as shown in the XPS
bonded to the quinone groups of the polydopamine not only by Schiff analysis (Fig. 6).
base reaction but also by Michael addition (Fig. 7a). The N 1 s spectra for
the membrane modified by both two-step and one-step methods show a 3.4. Fouling-reducing and self-cleaning capacities of the membranes
decrease in RNH2 intensity, which can be due to the condensation re­
action between PDA’s amino groups and enzymes’ carboxylic groups Oily wastewater filtration experiments were performed to determine
(Fig. 7a). Also, for the membrane modified by the two-step method, the potential of the immobilized lipase ET2 for fouling reduction and
there is a shift in the RNH2 binding energy, suggesting a change in the membrane cleaning by degrading the oil attached to the membrane
environment near the nitrogen. For the membrane modified by the one- surface. Membrane permeate flux was monitored during the filtration
step method, no shift was detected, indicating that PDA can cover part of experiments (Fig. 9b), and pure water permeance was evaluated by pure
the enzyme structure (Fig. 7b). water filtration before and after emulsion filtration, and after cleaning in

8
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Fig. 8. Water and n-heptane vapor adsorption at 23 ◦ C (left axis) and the ratio of water and n-heptane adsorption (right axis) for the pristine membrane, PDA-coated
membrane, and membranes with immobilized ET2 by simple adsorption, two-step, and one-step PDA coating. Note: The values are means of triplicates and their
respective standard deviations. Bars with different letters are significantly different according to Tukey’s test (p < 0.05).

100 mM SPB at pH 7 and 40 ◦ C (optimal conditions for ET2 activity) demonstrating that the one-step or one-pot ET2 immobilization with
(Fig. 9a). PDA can be used as an alternative to the conventional two-step pro­
Pure water filtration experiments showed that the PDA coating cedure. Therefore, simultaneous dopamine polymerization and ET2
reduced membrane permeance for water slightly (6 %), presumably due immobilization can be used to develop fouling-reducing enzymatic
to the penetration of PDA into membrane pores, which also decreased membranes using fewer chemicals in a shorter time.
the water vapor adsorption as shown in Fig. 8 and so, the water affinity Interestingly, the one-step ET2-immobilization method resulted in
of the membrane. Immobilization of ET2 on the PDA coating (i.e., the superior self-cleaning performance than the two-step method. As shown
two-step method), on the other hand, slightly increased membrane in Fig. 9a and Table 4, the pristine membrane recovered 48.0 ± 0.8 % of
permeance, which can be attributed to the increase in membrane surface its pure water permeance after the cleaning procedure, while the per­
hydrophilicity due to the presence of hydrophilic groups in the immo­ meance of the PDA-coated, two-step, and one-step ET2-immobilized
bilized enzyme structure. Although the membrane modified by the one- membranes resumed to 53.8 ± 1.4, 73.5 ± 0.7, and 89.7 ± 2.7 % of their
step ET2 coating strategy was the most hydrophilic (Fig. 8), it showed initial permeance, respectively. Therefore, besides decreasing mem­
initial water permeance similar to the PDA-coated membrane. This may brane fouling during emulsion filtration in a similar way as the two-step
be the result of the longer polymerization time used (12 h) compared approach, the enzyme immobilization by the one-step procedure had a
with that in the two-step method (6 h) and, hence, higher PDA loading. superior performance in terms of permeance recovery after cleaning,
During filtration of the oil emulsion, the oil content caused severe which can be a result of both increased water affinity (Fig. 8) and
fouling of the uncoated membrane, resulting in a 62 ± 2 % reduction in increased stability over reaction time as described in Fig. 3c.
membrane pure water permeance. The pristine membrane also showed The oil fouling control capacity (evaluated by pure water permeance
lower permeate flux during the oil emulsion filtration (Fig. 9b). The PDA reduction after filtration and backwash) of the membrane developed in
coating did not have a measurable impact on the pure water permeance this study was superior to those found in the literature for membranes
loss of the membrane after the two filtration cycles (Fig. 9a and Table 4); with immobilized lipases (Table 5). This can be attributed to several
however, it resulted in increased permeate flux during most of the oily factors, including the higher membrane hydrolytic activity, the cross-
wastewater filtration experiment, as it can be seen on Fig. 9b. The flow filtration mode (which contributes to decreasing the fouling)
immobilization of ET2 on the membrane surface, however, significantly instead of dead-end filtration, fewer filtration cycles, and also higher
reduced membrane fouling by the oil: pure water permeance reduction hydrophilicity since a ceramic membrane was used instead of a poly­
was limited to 40.2 ± 0.8 % and 35 ± 2 % for membranes modified by meric membrane. The self-cleaning capacity (evaluated by permeance
the two-step and one-step method, respectively. This may be attributed recovery after cleaning) was also superior to most of the studies, which
to the reduced adsorption of oil molecules on the ET2-immobilized can also be a consequence of the higher membrane hydrolytic activity.
membranes due to their higher hydrophilicity (Fig. 8) as well as the The membrane retention did not change considerably throughout the
hydrolysis of adsorbed oil molecules. The membrane with immobilized experiments. The pristine membrane had an oil retention of 90.6 ± 1.7
ET2 by the one-step approach also presented higher permeate fluxes %, the PDA coated membrane 87.8 ± 3.0 %, the two-step modified
throughout the experiment than the other membranes (Fig. 9b). The membrane 93.1 ± 2.8 %, and the one-step modified membrane 96.1 ±
better performance of the one-step modified membrane can be attrib­ 2.6 %. The same tendency of the water vapor adsorption (Fig. 8) can be
uted to the increased hydrophilicity (Fig. 8) and the enhanced stability noticed here, showing that the increased hydrophilicity of the modified
over reaction time (Fig. 3). membranes can also help to increase the membrane oil retention. Fig. S3
The results show that two-step and one-step immobilization led to shows a picture of the feed and the permeate of the one-step modified
similar pure water permeances (Fig. 9a) after the emulsion filtration, membrane.

9
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Fig. 9. Performance of pristine and enzyme-active α-alumina membranes in the filtration of soybean oil emulsion (1 g⋅L− 1): (a) water permeance before the emulsion
filtration (initial), after the final backwash (after filtration), and after the cleaning procedure; and (b) normalized permeate flux during the oil emulsion filtration.

As presented in Fig. 10, after three filtration and cleaning cycles, the
Table 4
pristine membrane had a reduction in pure water permeance of 91 %. In
Summary of pure water permeance changes for the emulsion filtration test as
contrast, the membrane with ET2 immobilized by the two-step method
shown in Fig. 9 for the pristine and enzymatically active membranes.
had a reduction of 69 % and the membrane modified by the one-step
Membrane Change in water permeance (%)
strategy had a reduction of only 38 %. After the first cycle, the pris­
Reduction after emulsion Recovery after tine, two-step and one-step modified membranes recovered 48 %, 73 %
filtration cleaning and 90 % of their initial pure water permeance, which decreased to 43
Pristine 61.5 ± 1.6 48.0 ± 0.8 %, 68 %, and 86 % after the second cycle, and 415, 61 %, and 81 % after
PDA 60.1 ± 1.7 53.8 ± 1.4 the third cycle, respectively. Even though decreasing over the filtration
PDA + ET2 (two- 40.2 ± 0.8 73.5 ± 0.7
and cleaning cycles, the permeance recovery showed by the membranes
step)
PDA + ET2 (one- 35.1 ± 2.0 89.7 ± 2.7 with immobilized ET2 was still significantly higher than the one shown
step) by the pristine membrane. The improved stability of the enzymes
immobilized by the one-step method (Fig. 3) results in increased reus­
ability over multiple filtration and cleaning cycles compared to the two-
step immobilized lipase. These results can also be a consequence of

10
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Table 5
Comparison of pure water permeance changes between this study and literature using membranes with immobilized lipase.
Membrane Lipase Immobilization MHA Feed Filtration Cleaning Changes in water Ref.
configuration and agent (µmol⋅min− 1⋅m− 2) conditions conditions permeance (%)
material
Reduction Recovery
after after
filtration cleaning

Tubular α-alumina ET2 PDA 1986 1 g⋅L− 1 Cross-flow, 1 SPB 100 35 90 This work
soybean oil bar, 1 L/min of mM pH 7,
in aqueous retentate flow 12 h, 40 ◦ C
SDS rate, 2 cycles of
solution 2 h each
Flat sheet Pancreatin1 EDC2 + NHS3 Not informed 0.5 g⋅L− 1 Dead-end, 6 PBS4 pH 8, 40 75 Shulze
polyethersulphone linseed oil cycles 12 h, 37 ◦ C et al.
(PES) in aqueous filtrating 400 (2017)
SDS mL
solution
Flat sheet PPL5 II and EDC + NHS 828 1 g⋅L− 1 Dead-end, 7–8 PBS pH 8, 58 72 Schmidt
polyvinylidene Candida rugosa linseed oil cycles 12 h, 37 ◦ C et al.
fluoride (PVDF) lipase in aqueous filtrating 200 (2018)
SDS mL
solution
Flat sheet PVDF Thermomyces Electron beam 823 1 g⋅L− 1 Dead-end, 4 PBS pH 8, 3 68 100 Schmidt
lanuginosus irradiation olive oil in cycles h, 37 ◦ C et al.
lipase aqueous filtrating 500 (2022)
SDS mL
solution
1
Mixture of lipase, protease and amylase.
2
1-ethyl-3-(3-dimethylaminopropyl) carbodiimide.
3
N-hydroxysuccinimid.
4
phosphate buffered saline.
5
porcine pancreatin lipase.

Fig. 10. Results of repeated filtration and cleaning experiments for the pristine, two-step and one-step modified membrane with PDA and ET2.

increased rigidity of the ET2 immobilized by the one-step approach methods rendered the membrane strong fouling resistance (reduction
since more enzyme moieties were possibly used for the immobilization. in pure water permeance of 35 % and 40 % after oil fouling, respec­
tively) and self-cleaning capability, with the one-step immobilization
4. Conclusions method outperforming the two-step method (90 % and 74 % of per­
meance recovery, respectively). Furthermore, the one-step enzyme
In this study, we demonstrated that one-step lipase ET2 immobili­ immobilization method using PDA is faster, uses fewer chemicals, con­
zation using polydopamine as the bonding agent was an excellent sumes less water, and produces less wastewater than the two-step
alternative to the two-step approach. Not only did it achieve an enzyme method, making it a lower-cost and more environmentally friendly
loading and a membrane hydrolytic activity similar to the conventional approach. For being a simpler method, the one-step immobilization can
two-step method, but it also enhanced the stability of the immobilized be easier to scale up, which, combined to the advantages described
enzyme. ET2 immobilization using both the one-step and two-step above, make it more attractive to industry and research institutions.

11
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

Declaration of Competing Interest improved industrial biocatalysts, Biotechnol. Adv. 36 (2018) 1470–1480, https://
doi.org/10.1016/j.biotechadv.2018.06.002.
[15] L. Zhou, X. Luo, J. Li, L. Ma, Y. He, Y. Jiang, L. Yin, L. Gao, Meso-molding three-
The authors declare that they have no known competing financial dimensionally ordered macroporous alumina: a new platform to immobilize
interests or personal relationships that could have appeared to influence enzymes with high performance, Biochem. Eng. J. 146 (2019) 60–68, https://doi.
the work reported in this paper. org/10.1016/j.bej.2019.03.002.
[16] N. Gao, W. Fan, Z.-K. Xu, Ceramic membrane with protein-resistant surface via
dopamine/diglycolamine co-deposition, Sep. Purif. Technol. 234 (2020), 116135,
Data availability https://doi.org/10.1016/j.seppur.2019.116135.
[17] X. Guo, C. Li, S. Fan, Z. Gao, L. Tong, H. Gao, Q. Zhou, H. Shao, Y. Liao, Q. Li,
W. Hu, Engineering polydopamine-glued sandwich-like nanocomposites with
Data will be made available on request. antifouling and antibacterial properties for the development of advanced mixed
matrix membranes, Sep. Purif. Technol. 237 (2020), 116326, https://doi.org/
10.1016/j.seppur.2019.116326.
Acknowledgments
[18] C. Chao, J. Liu, J. Wang, Y. Zhang, B. Zhang, Y. Zhang, X. Xiang, R. Chen, Surface
modification of halloysite nanotubes with dopamine for enzyme immobilization,
The authors thank the Coordenação de Aperfeiçoamento de Pessoal ACS Appl. Mater. Interfaces. 5 (2013) 10559–10564, https://doi.org/10.1021/
am4022973.
de Nível Superior – CAPES-PRINT Program (Project number
[19] H. Cheng, M. Hu, Q. Zhai, S. Li, Y. Jiang, Polydopamine tethered CPO/HRP-TiO2
88887.310560/2018-00) and Conselho Nacional de Desenvolvimento nano-composites with high bio-catalytic activity, stability and reusability: enzyme-
Científico e Tecnológico (CNPq, Brazil) for financial support, and the photo bifunctional synergistic catalysis in water treatment, Chem. Eng. J. 347
Laboratory of Membrane Processes (LABSEM/UFSC), Ceramic Process­ (2018) 703–710, https://doi.org/10.1016/j.cej.2018.04.083.
[20] W. Cheng, Y. Li, X. Li, W. Bai, Y. Liang, Preparation and characterization of PDA/
ing Laboratory (ProCer/UFSC), Supercritical Extraction and Thermo­ SiO2 nanofilm constructed macroporous monolith and its application in lipase
dynamics Laboratory (LATESC/UFSC), Dr. Li’s Lab (Rice University) and immobilization, J. Taiwan Inst. Chem. Eng. 104 (2019) 351–359, https://doi.org/
the Shared Equipment Authority at Rice University for making their 10.1016/j.jtice.2019.09.013.
[21] J. Mulinari, A. Ambrosi, M.D. de M. Innocentini, Y. Feng, Q. Li, M. Di Luccio, D.
research infrastructure available for this study. Hotza, J.V. Oliveira. Lipase immobilization on alumina membranes using a
traditional and a nature-inspired method for active degradation of oil fouling, Sep.
Purif. Technol. 287. (2022). 120527. https://doi.org/10.1016/j.
Appendix A. Supplementary data
seppur.2022.120527.
[22] Y. Ren, J.G. Rivera, L. He, H. Kulkarni, D.-K. Lee, P.B. Messersmith, Facile, high
Supplementary data to this article can be found online at https://doi. efficiency immobilization of lipase enzyme on magnetic iron oxide nanoparticles
org/10.1016/j.cej.2023.141516. via a biomimetic coating, BMC Biotechnol. 11 (2011) 63, https://doi.org/10.1186/
1472-6750-11-63.
[23] M. Schmidt, D. Breite, I. Thomas, M. Went, A. Prager, A. Schulze, Polymer
References membranes for active degradation of complex fouling mixtures, J. Membr. Sci. 563
(2018) 481–491, https://doi.org/10.1016/j.memsci.2018.06.013.
[24] A. Schulze, D. Breite, Y. Kim, M. Schmidt, I. Thomas, M. Went, K. Fischer,
[1] J. Chapman, A.E. Ismail, C.Z. Dinu, Industrial applications of enzymes: recent
A. Prager, Bio-inspired polymer membrane surface cleaning, Polymers. 9 (2017)
advances, techniques, and outlooks, Catalysts. 8 (2018) 238, https://doi.org/
97, https://doi.org/10.3390/polym9030097.
10.3390/catal8060238.
[25] A. Vanangamudi, D. Saeki, L.F. Dumée, M. Duke, T. Vasiljevic, H. Matsuyama,
[2] B.R. Facin, M.S. Melchiors, A. Valério, J.V. Oliveira, D. de Oliveira, Driving
X. Yang, Surface-engineered biocatalytic composite membranes for reduced
immobilized lipases as biocatalysts: 10 years state of the art and future prospects,
protein fouling and self-cleaning, ACS Appl. Mater. Interfaces. 10 (2018)
Ind. Eng. Chem. Res. 58 (2019) 5358–5378, https://doi.org/10.1021/acs.
27477–27487, https://doi.org/10.1021/acsami.8b07945.
iecr.9b00448.
[26] M. Schmidt, A. Prager, N. Schönherr, R. Gläser, A. Schulze, Reagent-free
[3] M. Aghababaie, M. Beheshti, A. Razmjou, A.-K. Bordbar, Covalent immobilization
immobilization of industrial lipases to develop lipolytic membranes with self-
of Candida rugosa lipase on a novel functionalized Fe3O4@SiO2 dip-coated
cleaning surfaces, Membranes. 12 (2022) 599, https://doi.org/10.3390/
nanocomposite membrane, Food Bioprod. Process. 100 (2016) 351–360, https://
membranes12060599.
doi.org/10.1016/j.fbp.2016.07.016.
[27] Y. Lan, D.W. Hiebner, E. Casey, Self-assembly and regeneration strategy for
[4] O. Barbosa, R. Torres, C. Ortiz, Á. Berenguer-Murcia, R.C. Rodrigues, R. Fernandez-
mitigation of membrane biofouling by the exploitation of enzymatic nanoparticles,
Lafuente, Heterofunctional supports in enzyme immobilization: from traditional
Chem. Eng. J. 412 (2021), 128666, https://doi.org/10.1016/j.cej.2021.128666.
immobilization protocols to opportunities in tuning enzyme properties,
[28] Z. Mehrabi, A. Taheri-Kafrani, M. Asadnia, A. Razmjou, Bienzymatic modification
Biomacromolecules. 14 (2013) 2433–2462, https://doi.org/10.1021/bm400762h.
of polymeric membranes to mitigate biofouling, Sep. Purif. Technol. 237 (2020),
[5] R. Fraas, M. Franzreb, Reversible covalent enzyme immobilization methods for
116464, https://doi.org/10.1016/j.seppur.2019.116464.
reuse of carriers, Biocatal. Biotransformation. 35 (2017) 337–348, https://doi.org/
[29] J. Lee, Y.-J. Won, D.-C. Choi, S. Lee, P.-K. Park, K.-H. Choo, H.-S. Oh, C.-H. Lee,
10.1080/10242422.2017.1344229.
Micro-patterned membranes with enzymatic quorum quenching activity to control
[6] A. Rodriguez-Abetxuko, D. Sánchez-deAlcázar, P. Muñumer, A. Beloqui, Tunable
biofouling in an MBR for wastewater treatment, J. Membr. Sci. 592 (2019),
polymeric scaffolds for enzyme immobilization, Front. Bioeng. Biotechnol. (2020),
117365, https://doi.org/10.1016/j.memsci.2019.117365.
https://doi.org/10.3389/fbioe.2020.00830.
[30] Z. Chen, Z. Sun, S. Ming, S. Li, Z. Zhu, W. Zhang, Bioinspired proteolytic membrane
[7] J. Zdarta, A.S. Meyer, T. Jesionowski, M. Pinelo, A general overview of support
(BPM) with bilayer pepsin structure for protein hydrolysis, Sep. Purif. Technol. 259
materials for enzyme immobilization: characteristics, Properties, Practical Utility,
(2021), 118214, https://doi.org/10.1016/j.seppur.2020.118214.
Catalysts. 8 (2018) 92, https://doi.org/10.3390/catal8020092.
[31] Y. Yurekli, Layer-by-layer self-assembly of multifunctional enzymatic UF
[8] S.B. Sigurdardóttir, J. Lehmann, S. Ovtar, J.-C. Grivel, M. Della Negra, A. Kaiser,
membranes, J. Appl. Polym. Sci. 137 (2020) 48750, https://doi.org/10.1002/
M. Pinelo, Enzyme immobilization on inorganic surfaces for membrane reactor
app.48750.
applications: mass transfer challenges, enzyme leakage and reuse of materials, Adv.
[32] W. Wang, X. Han, H. Yi, L. Zhang, The ultrafiltration efficiency and mechanism of
Synth. Catal. 360 (2018) 2578–2607, https://doi.org/10.1002/adsc.201800307.
transglutaminase enzymatic membrane reactor (EMR) for protein recovery from
[9] P. Zucca, E. Sanjust, Inorganic materials as supports for covalent enzyme
cheese whey, Int. Dairy J. 80 (2018) 52–61, https://doi.org/10.1016/j.
immobilization: methods and mechanisms, Molecules. 19 (2014) 14139–14194,
idairyj.2017.12.012.
https://doi.org/10.3390/molecules190914139.
[33] I. Kolesnyk, V. Konovalova, K. Kharchenko, A. Burban, J. Kujawa, W. Kujawski,
[10] L. Goldstein, G. Manecke. The chemistry of enzyme immobilization, in: L.B.
Enhanced transport and antifouling properties of polyethersulfone membranes
Wingard, E. Katchalski-Katzir, L. Goldstein (Eds.), Applied biochemistry and
modified with α-amylase incorporated in chitosan-based polymeric micelles,
bioengineering, Elsevier. 1976. pp. 23–126. https://doi.org/10.1016/B978-0-12-
J. Membr. Sci. 595 (2020), 117605, https://doi.org/10.1016/j.
041101-6.50008-X.
memsci.2019.117605.
[11] S. Chakraborty, H. Rusli, A. Nath, J. Sikder, C. Bhattacharjee, S. Curcio, E. Drioli,
[34] H. Zhang, Y. Wan, J. Luo, S.B. Darling, Drawing on membrane photocatalysis for
Immobilized biocatalytic process development and potential application in
fouling mitigation, ACS Appl. Mater. Interfaces. 13 (2021) 14844–14865, https://
membrane separation: a review, Crit. Rev. Biotechnol. 36 (2016) 43–58, https://
doi.org/10.1021/acsami.1c01131.
doi.org/10.3109/07388551.2014.923373.
[35] Y. Yang, H. Wang, J. Li, B. He, T. Wang, S. Liao, Novel functionalized nano-TiO2
[12] N.R. Mohamad, N.H.C. Marzuki, N.A. Buang, F. Huyop, R.A. Wahab, An overview
loading electrocatalytic membrane for oily wastewater treatment, Environ. Sci.
of technologies for immobilization of enzymes and surface analysis techniques for
Technol. 46 (2012) 6815–6821, https://doi.org/10.1021/es3000504.
immobilized enzymes, Biotechnol. Biotechnol. Equip. 29 (2015) 205–220, https://
[36] F. Chen, X. Shi, X. Chen, W. Chen, An iron (II) phthalocyanine/poly(vinylidene
doi.org/10.1080/13102818.2015.1008192.
fluoride) composite membrane with antifouling property and catalytic self-
[13] N. An, C.H. Zhou, X.Y. Zhuang, D.S. Tong, W.H. Yu, Immobilization of enzymes on
cleaning function for high-efficiency oil/water separation, J. Membr. Sci. 552
clay minerals for biocatalysts and biosensors, Appl. Clay Sci. 114 (2015) 283–296,
(2018) 295–304, https://doi.org/10.1016/j.memsci.2018.02.030.
https://doi.org/10.1016/j.clay.2015.05.029.
[14] C. Bernal, K. Rodríguez, R. Martínez, Integrating enzyme immobilization and
protein engineering: an alternative path for the development of novel and

12
J. Mulinari et al. Chemical Engineering Journal 459 (2023) 141516

[37] Y. Satyawali, K. Vanbroekhoven, W. Dejonghe, Process intensification: the future [46] T. Touqeer, M.W. Mumtaz, H. Mukhtar, A. Irfan, S. Akram, A. Shabbir, U. Rashid, I.
for enzymatic processes? Biochem. Eng. J. 121 (2017) 196–223, https://doi.org/ A. Nehdi, T.S.Y. Choong, Fe3O4-PDA-lipase as surface functionalized nano
10.1016/j.bej.2017.01.016. biocatalyst for the production of biodiesel using waste cooking oil as feedstock:
[38] Astm, C20 - Test methods for apparent porosity, water absorption, apparent characterization and process optimization, Energies. 13 (2019) 177, https://doi.
specific gravity, and bulk density of burned refractory brick and shapes by boiling org/10.3390/en13010177.
water, ASTM, International (2000), https://doi.org/10.1520/C0020-00R10. [47] A.M. Klibanov, Enzyme stabilization by immobilization, Anal. Biochem. 93 (1979)
[39] X. Fu, X. Zhu, K. Gao, J. Duan, Oil and fat hydrolysis with lipase from Aspergillus sp, 1–25, https://doi.org/10.1016/S0003-2697(79)80110-4.
JOACS. 72 (1995) 527–531. [48] D. Bresolin, B. Hawerroth, C. de Oliveira Romera, C. Sayer, P.H.H. de Araújo, D. de
[40] M.M. Bradford, A rapid and sensitive method for the quantitation of microgram Oliveira, Immobilization of lipase Eversa Transform 2.0 on poly(urea–urethane)
quantities of protein utilizing the principle of protein-dye binding, Anal. Biochem. nanoparticles obtained using a biopolyol from enzymatic glycerolysis, Bioprocess
72 (1976) 248–254, https://doi.org/10.1016/0003-2697(76)90527-3. Biosyst. Eng. 43 (2020) 1279–1286, https://doi.org/10.1007/s00449-020-02324-
[41] M.C. Proner, I. Ramalho Marques, A. Ambrosi, K. Rezzadori, C. da Costa, G. Zin, M. 6.
V. Tres, M. Di Luccio, Impact of MWCO and dopamine/polyethyleneimine [49] J. Kong, S. Yu, Fourier transform infrared spectroscopic analysis of protein
concentrations on surface properties and filtration performance of modified secondary structures, Acta Biochim. Biophys. Sin. (Shanghai) 39 (2007) 549–559,
membranes, Membranes. 10 (2020) 239, https://doi.org/10.3390/ https://doi.org/10.1111/j.1745-7270.2007.00320.x.
membranes10090239. [50] X. Cao, J. Luo, J.M. Woodley, Y. Wan, Mussel-inspired co-deposition to enhance
[42] N. Gao, Z.-K. Xu, Ceramic membranes with mussel-inspired and nanostructured bisphenol A removal in a bifacial enzymatic membrane reactor, Chem. Eng. J. 336
coatings for water-in-oil emulsions separation, Sep. Purif. Technol. 212 (2019) (2018) 315–324, https://doi.org/10.1016/j.cej.2017.12.042.
737–746, https://doi.org/10.1016/j.seppur.2018.11.084. [51] L. Dong, X. Liu, Z. Xiong, D. Sheng, C. Lin, Y. Zhou, Y. Yang, Fabrication of highly
[43] H. Treichel, M. Sbardelotto, B. Venturin, A.D. Agnol, J. Mulinari, S.M. Golunski, D. efficient ultraviolet absorbing PVDF membranes via surface polydopamine
B. Baldoni, C.B. Bevilacqua, R.J.S. Jacques, G.D.L.P.V. and A.J. Mossi,, Lipase deposition, J. Appl. Polym. Sci. 135 (2018) 45746, https://doi.org/10.1002/
production from a newly isolated Aspergillus niger by solid state fermentation using app.45746.
canola cake as substrate, Curr. Biotechnol. 5 (2016) 295–300, https://doi.org/ [52] M. Lu, J. Yu, in: Mussel-inspired Biomaterials for Cell and Tissue Engineering,
10.2174/2211550105666151124193225. Novel Biomaterials for Regenerative Medicine, Springer, Singapore, 2018,
[44] R.K. Nishihora, M.G.N. Quadri, D. Hotza, K. Rezwan, M. Wilhelm, Tape casting of pp. 451–474, https://doi.org/10.1007/978-981-13-0947-2_24.
polysiloxane-derived ceramic with controlled porosity and surface properties, [53] L.-S. Wang, S. Xu, S. Gopal, E. Kim, D. Kim, M. Brier, K. Solanki, J.S. Dordick, Facile
J. Eur. Ceram. Soc. 38 (2018) 4899–4905, https://doi.org/10.1016/j. fabrication of antibacterial and antiviral perhydrolase-polydopamine composite
jeurceramsoc.2018.07.016. coatings, Sci. Rep. 11 (2021) 12410, https://doi.org/10.1038/s41598-021-91925-
[45] T. Prenzel, K. Döge, R.P.O. Motta, M. Wilhelm, K. Rezwan, Controlled hierarchical 6.
porosity of hybrid ceramics by leaching water soluble templates and pyrolysis, [54] M. Kapoor, M.N. Gupta, Lipase promiscuity and its biochemical applications,
J. Eur. Ceram. Soc. 34 (2014) 1501–1509, https://doi.org/10.1016/j. Process Biochem. 47 (2012) 555–569, https://doi.org/10.1016/j.
jeurceramsoc.2013.11.033. procbio.2012.01.011.

13

You might also like