You are on page 1of 11

Process Biochemistry 51 (2016) 229–239

Contents lists available at ScienceDirect

Process Biochemistry
journal homepage: www.elsevier.com/locate/procbio

Immobilization of laccase via adsorption onto bimodal mesoporous


Zr-MOF
Shilong Pang a , Yanwen Wu b , Xiaoqiong Zhang c , Bingning Li b , Jie Ouyang a,∗ ,
Mingyu Ding c,∗∗
a
Department of Food Science and Engineering, College of Biological Sciences and Technology, Beijing Key Laboratory of Forest Food Process and Safety,
Beijing Forestry University, Beijing 100083, China
b
Beijing Center for Physical and Chemical Analysis, Beijing 100089, China
c
Department of Chemistry, Tsinghua University, Beijing 100084, China

a r t i c l e i n f o a b s t r a c t

Article history: A new bimodal micro-mesoporous Zr-metal organic framework (Zr-MOF, MMU) with a particle size
Received 30 August 2015 of approximately 200 nm was synthesized. The Brunauer–Emmett–Teller (BET) surface area and pore
Received in revised form 30 October 2015 diameter of the nanoscale MMU were 453.8 m2 /g and 3.5–7 nm, respectively. Laccase was immobi-
Accepted 30 November 2015
lized onto MMU via physical adsorption. This immobilized system exhibited a large adsorption capacity
Available online 2 December 2015
(221.83 mg/g), broad pH and temperature profiles, and better stability and repeatability than free laccase.
Inactivation of the interaction between the overcrowded laccase at high concentration and pore plugging
Keywords:
were not observed due to the secondary pore structure (3.5 nm). The immobilized laccase showed better
Metal organic framework
Immobilization
stability under extreme conditions than free laccase. The activity of immobilized laccase after utilization
Bimodal mesoporous 10 times remained at approximately 50%, and it remained at 55.4% of its initial activity at the end of 3
Laccase weeks of storage in an aqueous phase. The good stability and retention of enzymatic activity indicated
that the bimodal mesoporous Zr-MOF is a good support for the immobilization of laccase.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction basic studies focusing on the applications of laccase have been


conducted. However, the use of laccase in its native form is often
Enzymes are widely used in many types of industrial applica- hindered by several limitations, such as its sensitivity to the envi-
tions as highly efficient biological catalysts. However, the catalytic ronment, low operational stability and difficulties in recovery and
ability of enzymes is often hindered in harsh environments, such reuse.
as high-acid and high-alkaline solutions, due to transformations of Immobilization technology has been proven to be an effective
the enzyme structure. Laccase (EC1.10.3.2, p-benzenediol: oxygen method for implementing efficient and continuous applications of
oxidoreductase, size of 6.5 nm × 5.5 nm × 4.5 nm) [1], an oxidore- an enzyme [4]. Since the first immobilization experiment in 1916
ductase containing 500 amino acids (64–140 kDa, depending on [5], numerous materials have been applied to immobilize enzymes,
the type), is widely distributed in plants, fungi and some bacterial such as agarose, silica, TiO2 , organic gel, chitosan, kaolinite, multi-
strains. Laccase has a broad range of applications in the printing walled carbon nanotubes and graphene oxide [6]. Metal–organic
and dyeing industry, paper industry, and conversion of aromatic frameworks (MOFs) are compounds consisting of metal ions or
compounds [2] and removal of phenols, which leads to cancer and clusters coordinated to organic molecules (ligands) to form one-
teratogenicity due to waste water [3]. In recent years, exhaustive , two-, or three-dimensional structures. MOFs with large pores
and high specific surface areas have recently been synthesized,
particularly for applications in catalysis [7], drug delivery [8],
gas storage [9] and separation [10]. Different types of MOFs can
Abbreviations: MOF, metal organic framework; MMU, bimodal mesoporous Zr- easily be synthesized by changing the central metal ions and lig-
MOF; BET, Brunauer–Emmett–Teller; TMB, 1,3,5-trimethylbenzene; DMF, dimethyl ands [11], and these MOFs possess good Brunauer–Emmett–Teller
formamide; BSA, bovine serum albumin. (BET) surface areas and excellent dispersity compared with con-
∗ Corresponding author. Fax: +86 10 62338221.
∗∗ Corresponding author. Fax: +86 10 627970 87. ventional materials such as mesoporous silica materials, carbon
E-mail addresses: ouyangjie@bjfu.edu.cn (J. Ouyang), nanotubes and graphene. One of the attractive aspects of MOF
dingmy@mail.tsinghua.edu.cn (M. Ding). synthesis is that the organic bridging ligand can be synthetically

http://dx.doi.org/10.1016/j.procbio.2015.11.033
1359-5113/© 2015 Elsevier Ltd. All rights reserved.
230 S. Pang et al. / Process Biochemistry 51 (2016) 229–239

Fig. 1. Transmission electron microscopy (TEM) image of bimodal mesoporous Zr-MOF after activation.

modified to introduce a desired functionality to the framework, immobilizing laccase, and the quantity of adsorbed laccase, prop-
which can still possess good stability compared with conventional erties of immobilized laccase, reusability and storage stability were
materials such as chitosan and sodium alginate. For the above investigated.
reasons, MOFs can be considered a suitable carrier for biocata-
lyst immobilization. Most types of MOFs are microporous (pore 2. Materials and methods
size <2 nm). The small pore size prevents large molecules (such
as enzymes) from accessing the center sites inside MOFs. One 2.1. Materials
method for increasing the pore size is to expand the ligands:
a longer ligand can lead to either interpenetrating structures or Laccase (purified from white rot fungi) and ABTS (2,2 -azino-bis-
fragile frameworks that will collapse upon removal of the guests. (3-ethylbenzothiazoline-6-sulfonic acid)) were purchased from
Furthermore, it restricts the size of the pore aperture [12,13]. Thus, Sigma–Aldrich (St. Louis, MO, USA) and stored at 0–4 ◦ C before
the surfactant-templating method used to prepare various meso- use. Zirconium chloride (ZrCl4 ) was purchased from Strem Chem-
porous silicas can be applied to synthesize mesoporous MOFs by icals (USA). Terephthalic acid (BDC) was purchased from Tokyo
adding a chelating agent and surfactants to the reaction. The chelat- Chemical Industry Co., Ltd. (Japan). Bovine serum albumin (BSA),
ing agent can sufficiently establish the interaction between the Coomassie brilliant blue (G250), cetyltrimethylammonium bro-
framework building blocks and surfactants. The crystal growth is mide (CTAB), dimethyl formamide (DMF), 1,3,5-trimethylbenzene
thus directed by micelles at the nanoscale. Subsequently, crys- (TMB) and other chemicals were purchased from Beijing Chemicals
talline materials with carboxyl groups and two types of pores can be Corporation (Beijing, China).
achieved after removal of the guests [14,15].
There are several methods for enzyme immobilization, includ-
2.2. Preparation of the support
ing enzyme entrapment, absorption, encapsulation in an organic
or inorganic polymer, and covalent binding to the carrier or
MMU was synthesized in our laboratory using the hydrothermal
self-immobilization [16]. Each method has its own particular
method [20]. In detail, 0.7 g of ZrCl4 , 0.5 g of BDC, 0.324 g of CTAB,
strengths and weaknesses, e.g., entrapment, encapsulation and
30 mL of DMF, 4 mL of trifluoroacetic acid and 0.124 g of hydrochlo-
covalent binding have a relatively stronger interaction between
ric acid were mixed in a teflon tube. The mixture was sonicated
the enzyme and surface of the support, leading to a simultane-
for 10 min to form a homogeneous solution and then heated at
ous decrease in activity and enzyme loading. The first studies on
120 ◦ C for 24 h. Subsequently, the mixture was centrifuged, and
the immobilization of laccase date back to 2011 [17], but there are
the precipitated as-synthesized MMU was washed 3 times with
few reports on the immobilization of laccase using mesoporous
DMF to remove unreacted chemicals. To remove the template, the
MOFs other than the reports on microperoxidase-11 immo-
sediment was washed 3 times with 1 mol/L ammonium nitrate
bilized in Tb-mesoMOF [18] and zirconium- metalloporphyrin
(dissolved in ethanol/H2 O (1:2, v/v)) at 60 ◦ C for 24 h, followed by
as biomimetic catalysts [19].
washing with ethanol/H2 O (1:2, v/v). Finally, the mesostructured
In view of these problems, the present study primarily aimed to
MMU was desiccated at 60 ◦ C and activated at 190 ◦ C.
synthesize mesoporous Zr-MOFs (MMU), which are similar to UIO-
66 (UiO is the University of Oslo), using the surfactant-templating
method [20]. Zirconium was selected as the central metal ion rather 2.3. Characteristics of MMU
than copper because a series of HKUST (Hong Kong University of
Science and Technology) MOFs were easy to decompose in the Transmission electron microscopy (TEM) images of MMU were
aqueous phase and because the release of copper ions could deac- obtained using an H-7650B transmission electron microscope
tivate the adsorbed laccase. A previous study showed that UIO-66 (Hitachi, Japan) operating at 80 kV. N2 adsorption–desorption and
series materials possess excellent thermal stability, chemical stabil- pore size isotherms were obtained using a Micromeritics TriStar
ity and mechanical stability due to the use of competitive reagents 3020 II system (USA). The desorption branches of the isotherms
and high-symmetry subgroups [12]. Combined with the advantages were used to derive the pore size distribution based on the BJH
of UIO-66 and its own features, MMU was suitable for use in enzyme model, and BJH analyses were performed to evaluate the total
immobilization. In this study, MMU was used as the support for pore volume. Thermogravimetric analysis (TGA, Q50, USA) was per-
formed under a nitrogen atmosphere at a rate of 10 ◦ C/min up to
S. Pang et al. / Process Biochemistry 51 (2016) 229–239 231

Fig. 2. Nitrogen adsorption–desorption isotherms and distribution of pore diameters of bimodal mesoporous Zr-MOF after activation.

1000 ◦ C. Fourier transform infrared (FTIR) spectra were recorded The precipitate was washed 3 times with sodium acetate buffer
using a Perkin–Elmer spectrometer in the frequency S3 range (0.1 mol/L, pH 3.0) and dried at −20 ◦ C. The activities of the immo-
of 4000–500 cm−1 with a resolution of 4 cm−1 . X-ray diffraction bilized laccases with different adsorption concentrations were
(XRD) patterns were collected using a D8 Advance X-ray diffrac- detected. The activity recovery was also determined.
tometer (Bruker, Germany) with CuK␣ radiation ( = 1.5406 Å). The activity recovery of the immobilized laccase was calculated
Ultraviolet–visible (UV–vis) absorption spectra of proteins were using the following formula:
recorded using a U-3010 UV–vis spectrophotometer (Hitachi,
Japan). Ri
R (%) =
Rf
2.4. Adsorption of laccase on MMU where R is the activity recovery of the immobilized laccase (%), Ri
is the activity of immobilized laccase (IU/g), and Rf is the activity of
2.4.1. Assay of laccase activity free laccase (IU/g) under similar experimental conditions.
The activity of free laccase was determined spectrophotometri-
cally at 420 nm [21]. Briefly, 0.5 ␮mol/mL ABTS (2.9 mL, pH 3.0) and
10 mg/mL laccase (0.1 mL, pH 3.0) were mixed in a cuvette at 30 ◦ C 2.4.3. Optimization of the laccase adsorption time
for 20 min. A blank solution was used as a reference. The activity To evaluate the immobilization time of the support, approxi-
of immobilized laccase was determined by adding the immobi- mately 10 mg of activated MMU was immersed in 5 mL of laccase
lized laccase (approximately 10 mg) into 0.5 ␮mol/mL ABTS (pH solution (5 mg/mL dissolved in 0.1 mol/L sodium acetate buffer, pH
4.0, 20 mL) at 40 ◦ C. This mixture was centrifuged every 1.5 min, 3.0). The adsorption process was performed in a shaker at room
and the absorbance of 3 mL of the supernatant was measured at temperature for different times (0.5–3 h). The precipitate was sep-
420 nm. One unit (IU) of laccase activity was defined as the par- arated and washed 3 times with sodium acetate buffer (0.1 mol/L,
ticular specific variation (0.001) of the absorbance per minute. The pH 3.0) and then lyophilized. The activities of the immobilized
specific activity of the immobilized enzyme was calculated using laccases with different adsorption times were detected, and the
the following formula: activity recovery was calculated.

  A
SA IU/g = 2.4.4. Adsorption amount of the laccase
(M × m × t) When the adoption process finished, the adsorption amount
where SA is the specific activity (IU/g), A is the absorbance vari- was determined at 595 nm with a mixture containing 0.1 mL of
ation during a certain period of reaction time, M is the adsorption the supernatant, 0.9 mL of distilled water and 5 mL of G250, using
ability of the support (mg/g), m is the amount of support added to BSA as a reference (Bradford method). The standard curve was
the solution (g), and t is the reaction time (min). The support with- defined (y = 4.7017x − 0.0333, R2 = 0.9946) to calculate the laccase
out immobilized laccase was also tested. The results revealed that concentration in the supernatant. The adsorbed amount of MMU
MMU possessed no catalytic activity. was calculated using the following formula:
  (Y1 − Y2 )
2.4.2. Optimization of the laccase adsorption concentration M mg/g = 50 ×
m
To evaluate the loading capacity of the support, approxi-
mately 10 mg of activated MMU was immersed in 5 mL of sodium where M is the adsorption amount of the support (mg/g), Y1 is the
acetate buffer (0.1 mol/L, pH 3.0) with different initial laccase laccase concentration in the supernatant at a certain time, Y2 is the
concentrations (0.05–50 mg/mL, pH 3.0). The adsorption pro- laccase concentration in the supernatant at another time later than
cess was performed in a shaker at room temperature for 1 h. Y1 , and m is the quantity of support in the solution (g).
232 S. Pang et al. / Process Biochemistry 51 (2016) 229–239

2.5. Properties of immobilized laccase 3. Results and discussion

2.5.1. Effect of pH and temperature on the activities of free and 3.1. Characteristics of bimodal mesoporous Zr-MOF
immobilized laccases
A series of substrate solutions (0.5 ␮mol/mL ABTS) were pre- High-quality TEM images of MOFs are difficult to obtain because
pared with 0.1 mol/L sodium acetate at different pH values MOFs are typically sensitive to the electron beam. As shown in
(2.0–7.0). The catalytic activity was measured as described in Sec- Fig. 1(a), the particle size of MMU is approximately 200 nm, which
tion 2.4.1. is similar to that in a previous report [22]. The well-defined porous
The effect of pH on immobilized laccase was determined at room structure can be observed from the magnified view in Fig. 1(b),
temperature by adding the immobilized laccase (10 mg) to 20 mL which is consistent with other mesoporous MOFs reported by other
of the substrate at different pH values (2.0–7.0). The effect of pH on researchers [14,15]. This result preliminarily indicated that MMU
free laccase was determined by adding 0.1 mL of laccase solution was a mesoporous material. The hexagonal structure was not dis-
(10 mg/mL) to 2.9 mL of substrate at different pH values (2.0–7.0). tinct because of the small pore diameter. However, MMU retained
The effect of temperature on the activities of the immobilized the structure of UIO-66 [23] and a pore structure similar to a silica
and free laccases was determined in the temperature range from sphere because of the synthesis method of surfactant templating
20 to 70 ◦ C at pH 4.0 and 3.0 using preheated 0.5 ␮mol/mL ABTS [24].
(dissolved in 0.1 mol/L sodium acetate buffer, pH 3.0) as the sub- The BET surface area of MMU is relatively low (453.8 m2 /g)
strate. compared with other types of MOFs. However, it is similar to
Relative activity is defined as the specific activity of laccase at that of a silica sphere due to its synthesis method. The pore size
one point divided by the highest specific activity of laccase in the distribution from the BJH analysis displayed two main peaks of
same group of experiments. 3.5 and 7 nm, respectively, and the total pore volume of MMU
was 0.25 cm3 /g (Fig. 2). The N2 adsorption–desorption isotherms
showed two obvious hysteresis loops, demonstrating the existence
2.5.2. Effect of pH and temperature on the stabilities of free and of two pore systems. Similar hysteresis loops were also reported in
immobilized laccases previous studies [25]. The BET and BJH analytical models showed
The effect of pH on the stability of laccase was determined by that there were bimodal systems in MMU, one of which (smaller
incubating immobilized and free laccases in solutions with differ- one) was the pore of the framework building blocks and the other
ent pH values ranging from 2.0 to 7.0 for 1 h at room temperature, was formed by the framework building blocks, template agent, and
and the catalytic activity was measured under the optimal con- chelating agent. Compared with conventional UIO-66, the “micro-
ditions (free laccase: pH 3.0, 30 ◦ C; immobilized laccase: pH 4.0, porous” structure expanded to 3.5 nm because of the addition of
40 ◦ C), as described in Section 2.4.1. TMB. Recent studies regarding mesoporous MOFs are presented
The thermal stability was determined by incubating the immo- in Table 1 [26–30]. Among these studies, Cu3 (btc)2 (H2 O)3 was the
bilized and free laccases at different temperatures (20–70 ◦ C) for only MOF with a pore size sufficient to immobilize enzymes [29].
1 h. After cooling to room temperature, the enzyme activity was However, Cu3 (btc)2 (H2 O)3 could not be applied due to its poor
measured under the optimal conditions (free laccase: pH 3.0, 30 ◦ C; hydrological stability. Although the BET surface area of MMU was
immobilized laccase: pH 4.0, 40 ◦ C), as described in Section 2.4.1. lower than that of all other materials (Table 1), MMU with an ade-
quate pore diameter, stability in solution and carboxyl group to
help immobilize laccase was a suitable support for immobilization
2.5.3. Reusability and storage stability of immobilized laccase
compared with other types of MOFs.
The reusability of the immobilized laccase was assessed by per-
Fig. 3 presents the FTIR spectra of activated MMU with lac-
forming several consecutive operating cycles using 0.5 ␮mol/mL
case, activated MMU and UIO-66. The bands at 1500–1600 cm−1 ,
ABTS as the substrate. At the end of each cycle, the immobilized
which correspond to the C O of carboxylates, were coordinated
laccase was removed from the reaction medium and washed three
with metal centers by O during the deprotonation process. A
times with sodium acetate buffer (pH 4.0); the cycle was repeated
weak peak in the region of 1420–1480 cm−1 corresponded to the
with a fresh preheated aliquot of substrate mixed with the immo-
C C in the aromatic compound of the organic linker. The peak at
bilized laccase.
1667 cm−1 corresponded to DMF and disappeared after MMU was
To measure the storage stability, the immobilized laccase was
activated at 190 ◦ C. The strong peak at 1395 cm−1 was attributed to
stored at 4 ◦ C, and the remaining activity was assayed after storing
the C OH group of carboxylic acid. The bands observed in the region
for 3 weeks. The enzyme activity was measured under the optimal
of 750–655 cm−1 were assigned to the Zr O bond in the frame-
conditions (pH 4.0, 40 ◦ C), as described in Section 2.4.1.
work building blocks. The peaks at 1690 cm−1 (amide I band) and
1290 cm−1 (amide III band) in the spectrum of activated MMU with
2.6. Determination of kinetic parameters laccase confirmed that laccase was immobilized in MMU because
the amino group was not introduced during the synthesis process
The Michaelis–Menten kinetic parameters (Km ) and maximum and MMU with immobilized laccase was activated. The peak broad-
rate (Vmax ) of the free and immobilized laccases were determined ening and shifting of the spectrum of MMU with laccase indicated
by measuring the laccase activity with ABTS as substrate and initial that the internal environment was changed by electrostatic inter-
concentrations of 0.05–0.50 ␮mol/mL under the optimal conditions actions between laccase and the support.
(pH 4.0, 40 ◦ C), as described in Section 2.4.1. The data was calculated XRD patterns of UIO-66, activated MMU and MMU with laccase
according to Lineweaver–Burk plots. showed that activated MMU and MMU with laccase still presented
the two characteristic peaks (2 = 7.3◦ and 8.5◦ ) of UIO-66 [31,32],
which indicated that MMU retained the UIO-66 structure (Fig. 4).
2.7. Statistical analysis However, a significant reduction of the peaks in MMU with laccase
indicated that the crystal phase changed because of the immobi-
The experimental results were expressed as the mean value ± SD lization process, which also confirmed that laccase was successfully
of three parallel measurements. All statistical analyses were con- immobilized in MMU.
ducted using Microsoft Excel 2010.
S. Pang et al. / Process Biochemistry 51 (2016) 229–239 233

Table 1
Several recent results for mesoporous MOFs.

Compound Name Pore diameter (nm) BET (m2 /g) Ref.

Zn4 O(TPDC)3 Isoreticular metal–organic framework-16 2.9 – [26]


[Cd3 (bpdc)3 (dmf)]·5dmf·18H2 O Jilin University China-48 2.5 × 2.8 – [27]
Zn20 (cbIM)39 (OH) Zeolitic imidazolate frameworks-100 3.56 595 [28]
Tb16 (TATB)16 (DMA)24 Mesoporous MOF 1 3.9/4.7 1783 [29]
Cu3 (btc)2 (H2 O)3 Micro- and mesoporous MOFs 3.8–31.0 533–1225 [15]
Zr6 (O)4 (OH)10 (H2O)6 (TATB)2 Porous coordination network-777 3.8 2008 [12]
M2 (2,5-dioxidoterephthalate) M is Zn2+ , Mg2+ Isoreticular metal–organic framework-74-I to XI 1.4–9.8 1350–2510 [30]

Fig. 3. FTIR spectra of UIO-66 (bottom), bimodal mesoporous Zr-MOF after activation (middle) and bimodal mesoporous Zr-MOF with laccase (top). FTIR spectra were
recorded within the range 4000–500 cm−1 with a resolution of 4 cm−1 .

Fig. 4. XRD patterns of UIO-66 (top), bimodal mesoporous Zr-MOF after activation
(middle) and bimodal mesoporous Zr-MOF with laccase (bottom). Fig. 5. Thermogravimetric analysis (TGA) of bimodal mesoporous Zr-MOF (top) and
bimodal mesoporous Zr-MOF with laccase (bottom). Thermogravimetric analysis
was performed under a nitrogen atmosphere at a rate of 10 ◦ C/min up to 1000 ◦ C.

The thermogravimetric analysis curves are shown in Fig. 5. The 3.2. Optimization of the adsorption conditions for immobilized
mass loss in the first region of 50–100 ◦ C could be attributed to the laccase
desorption of adsorbed gas from the pore spaces. In the 150–250 ◦ C
region, mass loss occurred due to the removal of guest species 3.2.1. Optimization of the immobilization laccase concentration
(DMF) and carbonization of the immobilized laccase. The final mass The maximal activity of immobilized laccase was
loss in the region of 500–550 ◦ C was attributed to the disintegration 18316 ± 53 IU/g when the initial laccase concentration reached
of MOFs. The decomposition temperatures of MMU and MMU with 5 mg/mL. Meanwhile, the immobilized laccase exhibited maxi-
laccase confirmed that MMU possesses excellent thermostability mal activity recovery (95.90 ± 0.28%) and the amount of laccase
and that the immobilization process had no influence on its ther- adsorbed on MMU also reached the maximum (Fig. 6). The activ-
mostability [20,25]. The final residue amounts of MMU and MMU ity and adsorption amount of immobilized laccase increased
with laccase were 48.9% and 44.2%, respectively. with increasing initial adsorption concentration when the
234 S. Pang et al. / Process Biochemistry 51 (2016) 229–239

Fig. 6. Effect of the adsorption concentration on the immobilization of laccase on bimodal mesoporous Zr-MOF. The immobilization was performed in a shaker at room
temperature for 1 h at different initial concentrations. The immobilized laccase activity was assayed using 0.5 ␮mol/mL ABTS (20 mL) as a substrate at room temperature and
pH 3.0.

concentration was less than 5 mg/mL. The increasing amount of that immersing MMU in a high concentration of enzymatic solution
adsorbed laccase could increase the probability of contacting the for a long period of time did not lead to severe inactivation of the
substrate, and thus, the activity was improved. When the laccase laccase (87.06 ± 1.55% activity recovery remaining after adsorption
concentration reached 5–20 mg/mL, the activity recovery was for 3 h), although the pore diameter was not 2–4 times as large as
essentially constant (greater than 90%). In this case, the adsorption that of enzyme molecules [33]. This was due to the secondary pore
amount was almost unchanged, and the internal environment for size that allowed the substrate to enter; the relatively low BET sur-
adsorption and enzymatic reaction was not very crowded. A sig- face area, which made multi-point attachment between the laccase
nificant decrease in the activity recovery (remaining 75% activity and support more difficult; and the mass transfer resistance, which
of free laccase) was observed when the laccase concentration was made the laccase molecule difficult to desorb. Therefore, 1 h was
as high as 50 mg/mL, whereas the adsorbed amount of laccase chosen as the optimal immobilization time.
remained the same. This result could be explained by “competitive
adsorption” due to the high concentration of free laccase, which 3.2.3. Adsorption amount of laccase
was unable to completely enter the pores of the support. At the The highest amount of laccase adsorbed on MMU was
same time, laccase in the pores had to share a limited number of 221.83 ± 8.74 mg/g after incubation for 1 h in a 5 mg/mL laccase
carboxyl groups, representing an important way to be adsorbed solution, as shown in Figs. 6 and 7. The adsorption amount was
onto the support. The adsorption onto MMU was achieved by the higher than that of most of the supports applied in the immobiliza-
physical adsorption method, primarily relying on the hydrophilic, tion of laccase [25]. There was a slight decrease in the adsorption
electrostatic interactions and intermolecular forces between the amount (approximately 60 mg/g) over 1–3 h without a serious loss
enzyme and MMU. Therefore, the high concentration was not of activity (90% activity recovery remaining), as shown in in Fig. 7.
helpful in improving the activity due to the crowded internal After incubating for 4 days, the immobilized laccase was com-
environment. pletely separated from MMU, indicating that the desorption likely
led to the activity decrease in during the long immersion in the
3.2.2. Optimization of the immobilization time high concentration solution rather than the unexpected contrac-
MMU was immersed in a 5 mg/mL laccase (19100 ± 210 IU/g) tion between the crowed laccase molecules. To our knowledge, the
solution for different times (0.5–3 h). As shown in Fig. 7, an immo- excessive adsorbed concentration has three negative effects on the
bilization time of 1 h was considered to be ideal for reaching immobilized laccase activity [32]: (1) inducing the undesired lateral
the maximum immobilized activity (17928 ± 95 IU/g) and activity interaction between the over-crowded laccase molecules, (2) caus-
recovery (93.86 ± 0.49%). An increase in the immobilization time ing multi-point attachment between the laccase and support, and
(longer than 1 h) led to a decrease in the activity recovery and the (3) lowering the ABTS accessibility to immobilized laccase. How-
adsorption amount. This result could be explained by the physical ever, in this work, the pore diameter of the support was slightly
adsorption based on Fig. 7. The laccase adsorbed on the surface of larger than that of laccase molecules. This typical diameter could
the support was easily detached when it was immersed in solution prevent excessive laccase molecules from entering the channel to
for a long period of time. Consequently, the substrate was relatively form an undesired interaction. At the same time, the substrate and
difficult to contact the laccase inside the pores. Additionally, the degradation products can pass through the secondary pore size
desorbed laccase might affect the contact between laccase and the (3.5 nm), even though the 7-nm mesoporous material was blocked.
carboxyl group of the support, leading to a decrease in activity and Consequently, desorption was the main reason for the decrease
adsorption quantity, consistent with the adsorption amount results in activity, followed by the undesired interaction. In general, the
over 1–3 h in Fig. 7. The steady trend of the adsorbed amount after adsorption time varied (2–24 h), according to the different supports
2 h indicated that most of the laccase adsorbed on the surface of [32,34,35], and the smaller pore diameter increased the adsorp-
the support was desorbed, which also confirmed that most of the tion quantity but lowered the activity recovery [36,37]. Therefore,
laccase was adsorbed inside the pores. This result also confirmed a high laccase concentration was necessary to accelerate the
S. Pang et al. / Process Biochemistry 51 (2016) 229–239 235

Fig. 7. Effect of the adsorption time on the immobilization of laccase on bimodal mesoporous Zr-MOF. The immobilization was performed in a shaker at room temperature
and pH 3.0 for different lengths of time. The immobilized laccase activity was assayed using 0.5 ␮mol/mL ABTS (20 mL) as a substrate at room temperature and pH 3.0.

adsorption equilibrium and increase the possibility of laccase laccase, respectively. Fig. 9 shows that the activity of immobi-
molecules entering the pores. Moreover, there was no significant lized laccase tended to increase from 20 ◦ C until it reached the
decrease in the activity recovery when MMU was incubated in a maximum activity (15185 ± 502 IU/g) at 40 ◦ C, and then it quickly
relatively high concentration of laccase solution (Fig. 6). Therefore, decreased. By contrast, free laccase reached the highest activity
5 mg/mL and 1 h were chosen as the optimal adsorption concentra- (18250 ± 424 IU/g) at 30 ◦ C and declined dramatically above 30 ◦ C.
tion and adsorption time, respectively. The internal microenvironment and electrostatic force between the
immobilized laccase and support changed the structure of laccase,
3.3. Optimal conditions for laccase catalytic reaction which led to the change in the optimal temperature. It was clear
that the activities of immobilized laccase and free enzyme were
3.3.1. Optimal pH for laccase activity very close at low temperatures from 20 to 40 ◦ C and that the activity
The optimal pH for immobilized laccase activity depends on of immobilized laccase was higher than that of free laccase, indi-
the immobilization method and support used. As shown in Fig. 8, cating that immobilized laccase was more stable than free laccase.
the activity of free laccase reached the maximum at pH 3.0 This result may be due to the block effect of laccase immobiliza-
(14540 ± 594 IU/g) and then quickly decreased with increasing pH tion in the pore size of MMU [40]. The immobilized laccase retained
and was almost fully deactivated at the neutral condition. The activ- approximately 70% of its maximum activity at 70 ◦ C, indicating that
ity variation trend of immobilized laccase vs. pH was similar to it possessed good thermal stability.
that of free laccase. The relative activity of immobilized laccase
increased to approximately 90% when the pH changed from 2.0 to 3.4. Stability of immobilized laccase
3.0, and it reached the maximum level (12861 ± 768 IU/g) at pH 4.0.
Then, it decreased rapidly with increasing pH. Compared with free 3.4.1. Effect of pH on immobilized laccase stability
laccase, the optimal pH for immobilized laccase shifted from pH 3.0 In general, a high pH value will affect the stability of an immobi-
to 4.0 and showed a significantly higher activity than that of the free lized enzyme, including the process of adsorption and the enzyme
enzyme between pH 3.0 and 7.0, indicating that the immobilized catalytic reaction. Immobilized laccase had a broader range of
laccase was more stable than free laccase. applications than free laccase, but a significant decrease in the
Regarding free laccase, the ionic groups of laccase produced a activity of immobilized laccase at pH 7.0 was observed (Fig. 10). This
strong electrostatic repulsion when the pH value changed, leading result could also be explained by the desorption of laccase. The elec-
to stretching of the protein molecule, destruction and degeneration trostatic adsorption of the physical adsorption was achieved below
of the enzyme active center, and failure to decompose the substrate the isoelectric point (under the acidic condition). Under this con-
[38]. The decrease in the immobilized laccase activity with increas- dition, laccase was positively charged. Then, the positively charged
ing pH is due to the weakness of the electrostatic force between the laccase could be used to attract the negatively charged MMU
laccase and support. Compared with free laccase, the restriction of because of the carboxyl group on the surface. When the immo-
laccase in the pore site of MMU kept the internal microenvironment bilized laccase was in the alkaline condition, the electrostatic force
stable. Even at the extreme pH value of 7.0, the immobilized laccase was not significant, leading to desorption. The maximum activity
still retained 15% of its relative activity. The change in the opti- of immobilized laccase (17864 ± 118 IU/g) was slightly higher than
mal pH of immobilized laccase was because MMU was negatively that of free laccase (15100 ± 141 IU/g) after immersion for 1 h due to
charged; when laccase was adsorbed on the negatively charged the electrostatic force of the microstructures in negatively charged
support, the optimum pH moved to the alkaline range [39]. MMU changing the active structure of laccase [24].

3.3.2. Optimal temperature for laccase activity 3.4.2. Effect of temperature on immobilized laccase stability
According to the above result, the optimal pH values of 3.0 and The activities of free and immobilized laccase significantly
4.0 were used for the reaction of free laccase and immobilized changed in the temperature range from 20 to 70 ◦ C after 1 h of
236 S. Pang et al. / Process Biochemistry 51 (2016) 229–239

Fig. 8. Effect of pH on the activities of free laccase and immobilized laccase on bimodal mesoporous Zr-MOF. The free laccase (0.1 mL, 10 mg/mL) and immobilized laccase
(approximately 10 mg) activities were assayed using 0.5 ␮mol/mL ABTS (2.9 mL) and 0.5 ␮mol/mL ABTS (20 mL) as substrates, respectively, at room temperature to determine
the optimal pH value for the reaction.

Fig. 9. Effect of the reaction temperature on the activities of free laccase and immobilized laccase on bimodal mesoporous Zr-MOF. The free laccase (0.1 mL, 10 mg/mL)
and immobilized laccase (approximately 10 mg) activities were assayed using 0.5 ␮mol/mL ABTS (2.9 mL, pH 3.0) and 0.5 ␮mol/mL ABTS (20 mL, pH 4.0) as substrates to
determine the optimal reaction temperature.

incubation (Fig. 11). Clearly, the activities of the two types of lac- in the laccase conformation after adsorption onto MMU [46]. The
case were strongly dependent on temperature, and free laccase activity of immobilized laccase above 50 ◦ C decreased because a
exhibited the highest activity (5400 IU/g) at 30 ◦ C. Compared with large proportion of laccase on the MMU surface was inactivated due
free laccase without 1 h of incubation, the decrease in the maximal to the high temperature and immersion. Due to the high adsorption
activity was approximately 67% after the free laccase was incubated quantity and the mesoporous material, immobilized laccase exhib-
for 1 h. Some researchers have reported that laccase extracted from ited favorable resistance to thermal inactivation. If the temperature
Trametes versicolor retained 26.5% of its activity after 4 h of incu- could be controlled at approximately 40 ◦ C, the immobilized laccase
bation at 45 ◦ C [41], 3% after 2 h of incubation at 65 ◦ C [42], and would exhibit good usability.
0% after 1.5 h of incubation at 70 ◦ C [43]. Laccase extracted from
Rhus vernicifera retained 7% of its activity after 2 h of incubation
at 65 ◦ C [44]. Laccase extracted from white rot fungi retained 0% 3.5. Reusability and storage stability of immobilized laccase
of its activity after 4 h of incubation at 70 ◦ C [45]. These data indi-
cated that laccase was extremely sensitive to the environment and The reusability of immobilized laccase was studied because of its
easily deactivated after treatment at a relatively high temperature. importance in industrial applications to reduce costs. It was deter-
The activity of the immobilized laccase (12174 ± 155 IU/g) was 2.25 mined through 10 cycles at 40 ◦ C. The results indicated that the
times greater than that of free laccase after 1 h of incubation, similar activity of immobilized laccase decreased relatively quickly dur-
to free laccase without 1 h of incubation. This difference in activity ing the first 5 cycles, retaining 65.5% of its initial activity, and
was also due to a reduction in the molecular mobility and change was comparatively stable after 5 cycles (Fig. 12). This could be
explained by the denaturation of laccase on the surface of MMU
S. Pang et al. / Process Biochemistry 51 (2016) 229–239 237

Fig. 10. pH stabilities of free and immobilized laccases. The enzymes were incubated for 1 h under the optimal temperature and different pH values to investigate stability.
The free laccase activity was assayed at 30 ◦ C using 0.5 ␮mol/mL ABTS (2.9 mL, pH 3.0) as a substrate. The immobilized laccase activity was assayed at 40 ◦ C using 0.5 ␮mol/mL
ABTS (20 mL, pH 4.0) as a substrate.

Fig. 11. Temperature stabilities of free and immobilized laccases. The enzymes were incubated for 1 h under the optimal pH and different temperatures to investigate
stability. The free laccase activity was assayed at 30 ◦ C using 0.5 ␮mol/mL ABTS (2.9 mL, pH 3.0) as a substrate. The immobilized laccase activity was assayed at 40 ◦ C using
0.5 ␮mol/mL ABTS (20 mL, pH 4.0) as a substrate.

and the leakage of laccase from MMU during the oxidation reac- leaching of laccase from porous, such as in the case of activated
tion, in addition to the poisoning effect and diffusion limits caused sepharose and cellulose [51].
by the substrate [47]. Some studies have reported that immobilized Storage stability was tested by storing at 4 ◦ C for 3 weeks. The
laccase possessed better reusability [48,49]. The force of physi- results showed that immobilized laccase still retained 55.4% of its
cal adsorption between the laccase and MMU was weaker than initial activity for a period of 21 days. However, free laccase under
covalent attachment, leading to a low reusability. However, the the same storage conditions was completely inactivated, indicating
physical immobilization method had minimal negative impacts that immobilization could prevent laccase from inactivating and
on activity. Other researchers using the physical immobilization revealing the good storage stability.
method obtained reusability similar to that of the present study
[16,50]. Further improvement of the reusability of MMU is still
3.6. Kinetic parameters
necessary if considering the retaining rate. Covalent binding can
provide strong enzyme attachment. Therefore, carbodiimide acti-
Km is an important kinetic parameter in probing the bind-
vation can be used when the carboxylic group in the surface of MMU
ing ability of enzyme to substrate. The Km and Vmax reflect the
is expected to react with amino group of laccase for preventing the
effective characteristics of enzyme including both partition and
238 S. Pang et al. / Process Biochemistry 51 (2016) 229–239

Fig. 12. Operational stability of immobilized laccase. The immobilized laccase activity was assayed at 40 ◦ C using 0.5 ␮mol/mL ABTS (20 mL, pH 4.0) as a substrate. At the
end of each cycle, the immobilized laccase was removed from the reaction medium and washed three times with sodium acetate buffer (pH 4.0), and then the cycle was
repeated with a fresh preheated aliquot of substrate.

diffusion effects. In this study, the Km and Vmax of free enzyme were might arise from the relatively narrow pore size of MMU and the
0.330 ± 0.009 and 0.083 ± 0.002 mM/min, while those of immo- high initial adsorption concentration did not occur. The immo-
bilized laccase were 0.217 ± 0.005 and 0.072 ± 0.009 mM/min, bilized laccase exhibited higher activity over wider temperature
respectively. Low Km value indicates high affinity between enzyme and pH ranges than free laccase; in particular, the stability against
and substrate. The affinity of laccase with substrate increased heat denaturation increased significantly. The immobilized laccase
after immobilization, which was probably caused by structural exhibited good reusability and retained 55.45% of its initial activity
changes in the laccase introduced by immobilization procedure after storing at 4 ◦ C for 3 weeks. The Vmax /Km of laccase increased
[46]. Similar results have been reported recently [52,53]. The Vmax after immobilization and showed higher catalytic efficiency. The
value of immobilized laccase was 86.7% that of free laccase. The overall results show that, MMU can be used as a new support for
decrease in Vmax was in agreement with the previous reports immobilization and brings some advantages for biotechnological
that comparing free and immobilized laccase [53,54]. It might applications.
attribute to the substantial secondary structural perturbation and
changes on the enzyme surface upon adsorption [46,54]. As a con- Acknowledgements
sequence, the catalytic efficiency (Vmax /Km ) of immobilized laccase
(0.330 ± 0.049) was higher than that of free laccase (0.250 ± 0.002), The authors are thankful for the support of “the Fundamental
which proved that the immobilized laccase had relatively high Research Funds for the Central Universities (No. 2015ZCQ-SW-
catalytic efficiency. The catalytic efficiency was usually decreased 04)”, “Beijing Science and Technology Innovation Base Cultivation
upon immobilization in the similar reaction systems [52–54]. The and Development Projects (IG201307N)” and “Beijing Municipal
reason for the increase of the catalytic efficiency might be the Government Priorities and County Government Pretrigger Projects
bimodal mesoporous structure and the physical adsorption which (z121100000312010)”.
could help avoid enzyme/enzyme interaction and large aggregated
enzyme/support polymer layers.
Appendix A. Supplementary data

4. Conclusions
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.procbio.2015.11.
The success of an immobilization process primarily depends
033.
on the support used. In this work, a nanoscale bimodal (3.5
and 7 nm) mesoporous Zr-MOF modified with a carboxyl group
was synthesized using the surfactant-templating method to References
immobilize laccase. This material exhibited a high BET surface
[1] Q.H. Min, R.A. Wu, L. Zhao, H.Q. Qin, M.L. Ye, J.J. Zhu, Size-selective proteolysis
area (453.8 m2 /g), a bimodal mesoporous structure, hydrological on mesoporous silica-based trypsin nanoreactor for low-MW proteomo
stability and good thermal stability to resist 500 ◦ C. The immo- analysis, Chem. Commun. 46 (2010) 6144–6146.
bilization performed in a 5 mg/mL laccase solution for 1 h led to [2] E. Abadulla, T. Tzanov, S. Costa, K.H. Robra, A. Cavaco-Paulo, G.M. Gübitz,
Decolorization and detoxification of textile dye with a laccase from Trametes
the highest adsorption amount (221.83 ± 8.74 mg/g) and favor- hirsute, Appl. Environ. Microbiol. 66 (2000) 3357–3362.
able activity recovery (95.90 ± 0.28%). This excellent adsorption [3] L. Tang, G.M. Zeng, G.L. Shen, Y. Zhang, G.H. Huang, J. Li, Simultaneous
amount was better than that of most mesoporous materials. The amperometric determination of lignin peroxidase and manganese peroxidase
activities in compost bioremediation using artificial neural networks, Anal.
adsorption amount tended to be smooth and steady after 2 h Chim. Acta 579 (1) (2006) 109–116.
(175.00 ± 1.11 mg/g), proving that most of the laccase molecules [4] R.A. Sheldon, Enzyme immobilization: the quest for optimum performance,
were adsorbed inside the pores. Undesired lateral interactions that Adv. Synth. Catal. 349 (2007) 1289–1307.
S. Pang et al. / Process Biochemistry 51 (2016) 229–239 239

[5] M. Hartmann, X. Kostrov, Immobilization of enzymes on porous [30] H.X. Deng, S. Grunder, K.E. Cordova, C. Valente, H. Furukawa, M. Hmadeh,
silicas—benefits and challenges, Chem. Soc. Rev. 42 (2013) 6277–6289. et al., Large-pore apertures in a series of metal–organic frameworks, Science
[6] J.H. Park, H.H. Xue, J.S. Jung, K. Ryu, Immobilization of laccase on carbon 336 (2012) 1018–1023.
nanomaterials, Korean J. Chem. Eng. 29 (2012) 1409–1412. [31] C.L. Luu, T.T. Van Nguyen, T. Nguyen, T.C. Hoang, Synthesis, characterization
[7] O.K. Farha, A.M. Shultz, A.A. Sarjeant, S.T. Nguyen, J.T. Hupp, and adsorption ability of UiO-66-NH2 , Adv. Nat. Sci.: Nanosci. Nanotechnol. 6
Active-site-accessible, porphyrinic metal–organic framework materials, J. Am. (2015) 1–6.
Chem. Soc. 133 (2011) 5652–5655. [32] G.W. Petterson, J.B. DeCoste, G.T. Glover, Y. Huang, H. Jasuja, K.S. Walton,
[8] E.L. Kreno, K. Leong, O.K. Farha, M. Allendorf, R.P. Van Duyne, J.T. Hupp, Effects of pelletization pressure on the physical and chemical properties of
Metal–organic framework materials as chemical sensors, Chem. Rev. 112 the metal–organic frameworks Cu3 (BTC)2 and UiO-66, Microporous
(2012) 1105–1125. Mesoporous Mater. 179 (2013) 48–53.
[9] D. Britt, H. Furukawa, B. Wang, T.G. Glover, O.M. Yaghi, Highly efficient [33] N. Carlsson, H. Gustafsson, C. Thörn, L. Olsson, K. Holmberg, B. Akerman,
separation of carbon dioxide by a metal–organic framework replete with Enzymes immobilized in mesoporous silica: a physical–chemical perspective,
open metal sites, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 20637–20640. Adv. Colloid Interface Sci. 205 (2014) 339–360.
[10] J. An, S.J. Geib, N.L. Rosi, High and selective CO2 uptake in a cobalt adeninate [34] L.J. Dong, C.L. Ge, P.Y. Qin, Y. Chen, Q.H. Xu, Immobilization and catalytic
metal–organic framework exhibiting pyrimidine- and amino-decorated properties of candida lipolytic lipase on surface of organic intercalated and
pores, J. Am. Chem. Soc. 132 (2010) 38–39. modified MgAl-LDHs, Solid State Sci. 31 (2014) 8–15.
[11] J. Kim, B.L. Chen, T.M. Reineke, H.L. Li, M. Eddaoudi, D.B. Moler, Assembly of [35] J.A. Silva, G.A. Macedo, D.S.R. Gambetta, R.L.C. Giordano, L.R.B. Gonçalves,
metal–organic frameworks from large organic and inorganic secondary Immobilization of candida antarctica lipase B by covalent attachment on
building units: new examples and simplifying principles for complex chitosan-based hydrogels using different support activation strategies,
structures, J. Am. Chem. Soc. 123 (2001) 8239–8247. Biochem. Eng. J. 60 (2012) 16–24.
[12] D.W. Feng, K.C. Wang, J. Su, T.F. Liu, J. Park, Z.W. Wei, et al., A highly stable [36] Y.F. Zhu, S. Kaskel, J. Shi, T. Wage, K.H. van Pée, Immobilization of Trametes
zeotype mesoporous zirconium metal–organic-framework with ultralarge versicolor laccase on magnetically separable mesoporous silica spheres, Chem.
pores, Angew. Chem. Int. Ed. 54 (2015) 149–154. Mater. 19 (2007) 6408–6413.
[13] L.H. Wee, M.R. Lohe, N. Janssens, S. Kaskel, J.A. Martens, Fine tuning of the [37] F. Wang, C. Guo, L.R. Yang, C.Z. Liu, Magnetic mesoporous silica nanoparticles:
metal–organic framework Cu3 (BTC)2 KUST-1 crystal size in the 100 nm to fabrication and their laccase immobilization performance, Bioresour. Technol.
5 ␮m range, J. Mater. Chem. 22 (2012) 13742–13746. 101 (2010) 8931–8935.
[14] L.B. Sun, J.R. Li, J. Park, H.C. Zhou, Cooperative template-directed assembly of [38] N.M. Mubaraka, J.R. Wonga, K.W. Tan, J.N. Sahu, E.C. Abdullah, N.S. Jayakumar,
mesoporous metal–organic frameworks, J. Am. Chem. Soc. 134 (2012) et al., Immobilization of cellulose enzyme on functionalized multiwall carbon
126–129. nanotubes, J. Mol. Catal. B: Enzym. 107 (2014) 124–131.
[15] L.G. Qiu, T. Xu, Z.Q. Li, W. Wang, Y. Wu, X. Jiang, et al., Hierarchically micro- [39] T.C. Hung, N. Raghavan, G.R. Nair, S.H. Chiou, W.T. Wu, Binary immobilization
and mesoporous metal–organic frameworks with tunable porosity, Angew. of Candida rugosa lipase on chitosan, J. Mol. Catal. B: Enzym. 26 (2003) 69–78.
Chem. Int. Ed. 47 (2008) 9487–9491. [40] F. Wang, C. Guo, H.Z. Liu, C.Z. Liu, Immobilization of Pycnoporus sanguineus
[16] M. Fernández-Fernández, M.Á. Sanromán, D. Moldes, Recent developments laccase by metal affinity adsorption on magnetic chelator particles, J. Chem.
and applications of immobilized laccase, Biotechnol. Adv. 31 (2012) Technol. Biotechnol. l83 (2008) 97–104.
1808–1825. [41] M.E. Çorman, N. Öztürk, N. Bereli, S. Akgöl, A. Denizlib, Preparation of
[17] Z. Zhou, M. Hartmann, Progress in enzyme immobilization in ordered nanoparticles which contains histidine for immobilization of Trametes
mesoporous materials and related applications, Chem. Soc. Rev. 42 (2013) versicolor laccase, J. Mol. Catal. B: Enzym. 63 (2010) 102–107.
3894. [42] G. Bayramoglu, M. Yilmaz, M.Y. Arica, Reversible immobilization of laccase to
[18] V. Lykourinou, Y. Chen, S. Wang, L. Meng, T. Hoang, L.J. Ming, Immobilization poly (4-vinylpyridine) grafted and Cu(II) chelated magnetic beads:
of MP-11 into a mesoporous metal–organic framework, MP-11@ mesoMOF—a biodegradation of reactive dyes, Bioresour. Technol. 101 (2010) 6615–6621.
new platform for enzymatic catalysis, J. Am. Chem. Soc. 133 (2011) [43] A. Pich, S. Bhattacharya, H.P. Adler, T. Wage, A. Taubenberger, Z. Li, et al.,
10382–10385. Composite magnetic particles as carriers for laccase from Trametes versicolor,
[19] D.W. Feng, Z.Y. Gu, J.R. Li, H.L. Jiang, Z.W. Wei, H.C. Zhou, Macromol. Biosci. 6 (2006) 301–310.
Zirconium-metalloporphyrin PCN-222: mesoporous metal–organic [44] M.Y. Arica, B. Altintas, G. Bayramoglu, Immobilization of laccase onto
frameworks with ultrahigh stability as biomimetic, Angew. Chem. Int. Ed. 124 spacer-arm attached non-porous poly (GMA/EGDMA) beads: application for
(2012) 10453–10456. textile dye degradation, Bioresour. Technol. 100 (2009) 665–669.
[20] J. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, B. Silvia, A new [45] L.F. Qiu, Z.X. Huang, The treatment of chlorophenols with laccase immobilized
zirconium inorganic building brick forming metalorganic frameworks with on sol–gel-derived silica, World J. Microbiol. Biotechnol. 26 (2010) 775–781.
exceptional stability, J. Am. Chem. Soc. 130 (2008) 13850–13851. [46] N. Tüzmen, T. Kalburcu, A. Denizli, Immobilization of catalase via adsorption
[21] R. Bourbonnais, M.G. Paice, Oxidation of nonphenolic substrates—an onto metal–chelated affinity cryogels, Process Biochem. 47 (2012) 26–33.
expanded role for laccase in lignin biodegradation, FEBS Lett. 267 (1990) [47] M. Sari, S. Akgöl, M. Karataş, A. Denizli, Reversible immobilization of catalase
99–102. by metal chelate affinity interaction on magnetic beads, Ind. Eng. Chem. Res.
[22] H.R. Abid, H.Y. Tian, H.M. Ang, M.O. Tade, C.E. Buchley, S.B. Wang, Nanosize 45 (2006) 3036–3043.
Zr-metal organic framework (UiO-66) for hydrogen and carbon dioxide [48] Y.G. Makas, N.A. Kalkan, S. Aksoy, H. Altinok, N. Hasirci, Immobilization of
storage, Chem. Eng. J. 187 (2012) 415–420. laccase in ␬-carrageenan based semi-interpenetrating polymer networks, J.
[23] F. Vermoortele, G.L. Bars, B. Bueken, B. Van de Voorde, M. Vandichel, K. Biotechnol. 148 (2010) 216–220.
Houthoofd, et al., Synthesis modulation as a tool to increase the catalytic [49] H.J. Qiu, C.X. Xu, X.R. Huang, Y. Ding, Y.B. Qu, P.J. Gao, Adsorption of laccase on
activity of MOFs: the unique case of UiO-66 (Zr), J. Am. Chem. Soc. 135 (2013) the surface of nanoporous gold and the direct electron transfer between
11465–11468. them, J. Phys. Chem. C 112 (38) (2008) 14781–14785.
[24] D.Y. Zhao, Q.S. Huo, J.L. Feng, B.F.S. Chmelka, G.D. tucky, Nonionic triblock and [50] Y. Wang, X.H. Zheng, M. Zhao, Study of immobilization of laccase on
star diblock copolymer and oligomeric surfactant syntheses of highly ordered, mesoporous molecular sieve MCM-41, J. Chem. Eng. Chin. Univ. 22 (2008)
hydrothermally stable, mesoporous silica structures, J. Am. Chem. Soc. 120 83–87.
(1998) 6024–6036. [51] N. Durán, M.A. Rosa, A. D’Annibale, L. Gianfreda, Applications of laccases and
[25] Y.Y. Liu, Z.T. Zeng, G.M. Zeng, L. Tang, Y. Pang, Z. Li, et al., Immobilization of tyrosinases (phenoloxidases) immobilized on different supports: a review,
laccase on magnetic bimodal mesoporous carbon and the application in the Enzyme Microb. Technol. 31 (2002) 907–931.
removal of phenolic compounds, Bioresour. Technol. 115 (2011) 21–26. [52] H. Sun, H. Yang, W. Huang, S. Zhang, Immobilization of laccase in a
[26] M. Eddaoudi, J. Kim, N.L. Rosi, D. Vodak, J. Wachter, M. O’Keeffe, et al., sponge-like hydrogel for enhanced durability in enzymatic degradation of dye
Systematic design of pore size and functionality in isoreticular MOFs and their pollutants, J. Colloid Interface Sci. 450 (2015) 353–360.
application in methane storage, Science 295 (2002) 469–472. [53] M. Deng, H. Zhao, S. Zhang, C. Tian, Z. Di, P. Du, et al., High catalytic activity of
[27] Q. Fang, G. Zhu, Z. Jin, Y. Ji, J. Ye, M. Xue, et al., Mesoporous metal–organic immobilized laccase on core–shell magnetic nanoparticles by dopamine
framework with rare etb topology for hydrogen storage and dye assembly, self-polymerization, J. Mol. Catal. B: Enzym. 112 (2015) 15–24.
Angew. Chem. Int. Ed. 46 (2007) 6638–6642. [54] A.P.M. Tavares, C.G. Silva, G. Dražić, A.M.T. Silva, J.M. Loureiro, L.F. Joaquim,
[28] B. Wang, A.P. Côté, H. Furukawa, M. O’Keeffe, O.M. Yaghi, Colossal cages in Laccase immobilization over multi-walled carbon nanotubes: kinetic,
zeolitic imidazolate frameworks as selective carbon dioxide reservoirs, thermodynamic and stability studies, J. Colloid Interface Sci. 454 (2015)
Nature 453 (2008) 207–211. 52–60.
[29] Y.K. Park, S.B. Choi, H. Kim, K. Kim, B. Won, K. Choi, et al., Crystal structure and
guest uptake of a mesoporous metal–organic framework containing cages of
3.9 and 4.7 nm in diameter, Angew. Chem. Int. Ed. 46 (2007) 8230–8233.

You might also like