Interaction GO and Tyr residues in alfa amilase

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Hazardous Materials 453 (2023) 131389

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Research Paper

Intrinsic mechanisms for the inhibition effect of graphene oxide on the


catalysis activity of alpha amylase
Xinwei Liu 1, Binbin Sun 1, Chunyi Xu , Tianxu Zhang , Yinqing Zhang , Lingyan Zhu *
Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), Tianjin Key Laboratory of Environmental Remediation and Pollution Control,
College of Environmental Science and Engineering, Nankai University, Tianjin 300350, PR China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Binding interaction between GO and


α-amylase was firstly investigated.
• GO strongly binds with α-amylase in a
spontaneous and exothermic manner.
• GO binds with α-amylase mainly via
hydrogen bonding and van der Waals
force.
• GO quenches intrinsic fluorescence of
α-amylase by changing secondary
conformation.
• GO reduces the ability of α-amylase to
catalyze digestion of starch into glucose.

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

Lingxin CHEN Comprehending the interactions between graphene oxide (GO) and enzymes is critical for understanding the
toxicities of GO. In this study, the inherent interactions of GO with α-amylase as a typical enzyme, and the
Keywords: impacts of GO on the conformation and biological activities of α-amylase were systematically investigated. The
Graphene oxide results reveal that GO formed ground-state complex with α-amylase primarily via hydrogen bonding and van der
α-amylase
Waals interactions, thus quenching the intrinsic fluorescence of the protein statically. Particularly, the strong
Binding interaction
interactions altered the microenvironment of tyrosine and tryptophan residues, caused rearrangement of poly­
Conformational structure
Biological activity peptide structure, and reduced the contents of α-helices and β-sheets, thus changing the conformational structure
of α-amylase. According to molecular docking results, GO binds with the amino acid residues (i.e., His299,
Asp300, and His305) of α-amylase mainly through hydrogen bonding, which is in accordance with in vitro in­
cubation experiments. As a consequence, the ability of α-amylase to catalyze starch hydrolysis into glucose was
depressed by GO, suggesting that GO might cause dysfunction of α-amylase. This study discloses the intrinsic
binding mechanisms of GO with α-amylase and provides novel insights into the adverse effects of GO as it enters
organisms.

* Corresponding author.
E-mail address: zhuly@nankai.edu.cn (L. Zhu).
1
These authors contributed equally to this work and should be considered co-first authors.

https://doi.org/10.1016/j.jhazmat.2023.131389
Received 11 January 2023; Received in revised form 25 March 2023; Accepted 7 April 2023
Available online 8 April 2023
0304-3894/© 2023 Elsevier B.V. All rights reserved.
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

1. Introduction underneath the active site and assists the reaction, while the enzyme
structure is stabilized by a calcium ion. Any increase or decrease in
Because of the distinctive properties, such as large surface area, good enzymatic activity can result in a wide range of human diseases and
biocompatibilities, and strong affinities with biomacromolecules (e.g., disorders, such as metabolic imbalance and specific enzyme mutations
enzymes), graphene oxide (GO), a kind of typical and common two- [26]. α-amylase is one of the most important digestive enzymes in vivo
dimensional (2D) nanomaterial, finds extensive applications in and the inhibitory effects on α-amylase activity may result in appetite
biomedical and environmental fields [1,2]. GO has found many appli­ loss and glucose metabolism disorders, which can lead to abnormalities
cations in biomedical fields, such as drug and protein delivery, bio­ in body function [27]. Dhobale et al. discovered that ZnO nanoparticles
sensors and bioimaging, and antimicrobial agents [3–5]. GO has higher at 20 mg/L inhibited α-amylase activity by 49% [28]. Jiang et al. found
loading efficiency and biocompatibility with drug molecules or/and that spherical and polygonal starch nanoparticles inhibited the enzy­
biomolecules and better stability in physiological environment than matic activity of α-amylase by impacting its spatial conformation [29].
other therapeutic agents, such as metal nanoparticles, polymers, and In contrast, citrate-stabilized gold nanoparticles and silver nanoparticles
other carbon-based nanomaterials [6–8]. Large-scale applications and significantly enhanced the specific activity of α-amylase [30,31]. As one
improper discharges may lead to the inevitability of GO release into the typical 2D nanomaterial with plentiful functional groups, it is hypoth­
natural aquatic environment [9]. As a result, it may pose direct or in­ esized that GO might interact with α-amylase actively and change its
direct adverse effects to aquatic organisms and eventually human beings conformational structure and biological activity differently from other
through food chains [10,11]. Once GO enters organisms through mul­ nanoparticles.
tiple exposure pathways (e.g., oral ingestion, inhalation, or skin con­ Herein, the interactions of GO and α-amylase at molecular level were
tact), it may firstly contact with biomacromolecules widely existing in systematically explored through various spectroscopic characterizations
biofluids. Understanding the interactions between GO and bio­ and theoretical calculations. Enzymatic activities of α-amylase in the
macromolecules at molecular level is of importance for better predicting absence and presence of GO were also investigated. This work would
its biological effects. Current studies on the interactions between GO and provide novel insights into affinities of enzymes with GO and the change
biomacromolecules mainly focus on serum albumin, fibrinogen, and trend of conformational structure and biological effects induced by GO.
immunoglobulin, and the results reveal that GO may significantly
change the spatial conformation of proteins and finally influence their 2. Materials and methods
physiological functions [12,13]. As another important component of
serum biomacromolecules, enzymes show different conformational 2.1. Materials
structure and biological functions, while their interactions with GO have
been largely overlooked. GO powder was purchased from Xianfeng Nano Materials Technol­
As typical and widely existing biological catalysts, enzymes partici­ ogies Co., Ltd (Nanjing, China). Lyophilized powder of α-amylase
pate in specific metabolic reactions, such as energy transformation and (10065, ≥98%) was provided by Sigma Aldrich company (St. Louis, MO,
bond hydrolysis, while their structures and activities remain unchanged USA). Thioflavin T (ThT, D844310, ≥97%) was supplied by Macklin
before and after reactions [14,15]. Previous studies found that GO had Biochemical Co., Ltd (Shanghai, China). Phosphate-buffered saline (PBS,
greater adsorption capacity for enzymes (e.g., lysozyme) than other 10 mmol/L, pH 7.4) was obtained through dissolving 2.9 g Na2HPO4,
common nanoparticles like single-walled and multi-walled carbon 0.2 g KH2PO4, 8.0 g NaCl, and 0.2 g KCl into 1 L ultrapure water.
nanotubes, which may be due to the strong hydrophobic and electro­ α-amylase was dissolved in the PBS solution and stored at 4 ◦ C. Ultra­
static interactions of GO with enzymes [16,17]. The interactions be­ pure water (18.2 MΩ/cm, HHitech, China) was utilized throughout the
tween GO and enzymes might alter the conformational structure and experimental studies.
catalytic functions of enzymes. For example, Huang et al. discovered
that GO bound to the allosteric sites of trypsin and inhibited its activities 2.2. GO preparation and characterization
in a non-competitive way [18]. Conversely, the peroxidase activity of
cytochrome c was improved after binding to GO since the active sites The GO stock solution was prepared according to our earlier research
were slightly perturbed [19]. These discrepant results suggest that the with a few modifications [32]. Briefly, 80 mg of GO powder was mixed
impacts of GO on enzymes’ catalytic activities and conformational with 400 mL of ultrapure water and stirred for 12 h at 300 rpm/min. The
change are related to the physicochemical properties of enzymes, such GO suspension was ultrasonically treated in an ice bath for 6 h to obtain
as surface charge, amino acid amount, molecular weight, and active a uniform stock solution. Before each set of experiment, the GO stock
sites. solution was ultrasonicated for 15 min. Field emission scanning electron
Previous research discovered that after GO treatment in the parent microscopy (FE-SEM, Gemini 500, ZEISS, Germany) and high-resolution
zebrafish, the biosynthesis and metabolism of glucose in their offspring transmission electron microscopy (HR-TEM, 2100F, JEOL, Japan) were
zebrafish were significantly disturbed [20]. GO also downregulated the used to characterize the morphological feature and lateral size of GO.
carbohydrate metabolism in pepper fruit and markedly increased the The surface chemistry and elemental compositions of GO were deter­
number of starch grains in single-celled organism Chlorella vulgaris [21, mined using X-ray photoelectron spectrometer (XPS, PHI 5000 VersaP­
22]. All these metabolic processes are associated with α-amylase, which robe, ULVAC-PHI, Japan), FT-IR spectrometer (TENSOR 37, Bruker,
is a kind of hydrolytic enzyme that catalyzes the digestion of starch into Germany), elemental analyzer (vario EL cube, elementar, Germany),
monosaccharides or oligosaccharides and finally influences the carbo­ Raman spectra (Renishaw inVia, UK), and UV-Vis spectrometer
hydrate metabolism [23]. α-amylase is widely distributed in animals (e. (UV-2600, Shimadzu, Japan). Our previous study provided a compre­
g., salivary gland, pancreas, and blood), plants, and microorganisms hensive description of the characterization methods [33].
[24]. The molecular weight of pancreatic α-amylase in human body is
about 57.78 kDa, containing 511 amino acid residues in one single 2.3. Interaction between GO and α-amylase
oligosaccharide chain, and the detailed physicochemical properties are
shown in Table S1 [25]. The basic structure of α-amylase consists of 2.3.1. Adsorption experiments
three structural domains (A, B, and C, respectively). The largest domain Batch experiments were conducted at room temperature to reveal the
A is a barrel of eight parallel stretches of extended chain enclosed by adsorption affinity of α-amylase (from 50 to 1000 mg/L) on GO (100
eight helices [24]. α-amylase cleaves the link between the two sugars in mg/L). A rotary shaking incubator was used to shake the vials for 12 h,
the starch chain with the help of three acidic groups (i.e., glutamate 233, and all experiments were carried out in triplicate. After incubation, the
aspartate 197, and aspartate 300) at its active site. A chloride ion binds suspensions were centrifuged at 8000 rpm/min for 15 min before being

2
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

filtered through 0.22 µm nylon membrane. The filtered protein sus­ 2.3.6. Circular dichroism (CD)
pension was placed in the microplate (NEST Biotechnology, China) and The far-UV CD spectra of α-amylase and GO complexes were ob­
the Bradford protein assay kit was used to determine its concentration tained by Spectrometer J-810 (Jasco, Japan) using a quartz cell with a
[34]. Langmuir and Freundlich models were utilized to fit the adsorption path length of 0.1 cm. An accumulation of three scans with 200 nm/min
isotherms to understand the adsorption mechanism. speed was performed and spectra were obtained from 185 to 260 nm
with a band width of 2 nm. α-amylase concentration was fixed at 3000
2.3.2. Size and zeta potential measurements mg/L, while that of GO varied between 0 and 75 mg/L. Appropriate
The hydrodynamic diameter (Dh) and zeta potential of GO before and background corrections were applied to all spectra.
after incubation with α-amylase in PBS were determined by Malvern
Zetasizer (Nano ZS90, Malvern, UK). Zeta potential would be automat­ 2.3.7. Fluorescent dying experiment with ThT
ically obtained by the conversion of electrophoretic mobilities using the The conformational information of α-amylase was characterized
zeta potential analyzer of Malvern Zetasizer. using the fluorescent dye ThT [37]. α-amylase at 300 mg/L was firstly
incubated with various concentrations of GO (from 0 to 100 mg/L) for
2.3.3. Fluorescence spectroscopy 0.5 h; 120 μM ThT was then added into the suspensions and further
The fluorescence intensity changes of α-amylase caused by GO of incubated for another 0.5 h. The fluorescence spectra of α-amylase-ThT
various concentrations (0–100 mg/L) were recorded at different incu­ conjugates with and without GO were measured with an excitation
bation time (0, 30, and 60 min) and temperatures (277 K, 298 K, and wavelength of 440 nm and slit width of 5 nm in the range from 400 to
310 K). Fluorescence emission spectra of α-amylase as a function of GO 600 nm.
concentration ranging from 290 to 500 nm were recorded by a fluo­
rescence spectrophotometer installed with Xe lamp (F-7000, Hitachi, 2.3.8. Assay of α-amylase activity
Japan) with the excitation wavelength of 280 nm and slit width of 2.5 The starch hydrolytic activity of α-amylase with and without GO was
nm. evaluated by monitoring the sugar release utilizing the 3, 5-dinitrosali­
Time-resolved fluorescence spectra of α-amylase and α-amylase-GO cylic acid (DNS) method in accordance with the procedure reported by
system were recorded on a FLS1000 high sensitivity photoluminescence Bernfeld et al. [38]. The α-amylase assay kits (Cat#BC0610) were ob­
spectrometer with both steady state and time-resolved mode (Edin­ tained from Solarbio Science & Technology Co., Ltd (Beijing, China).
burgh, England) at 298 K. Fluorescence decay trace was recorded via the The α-amylase-GO complex was prepared by mixing 250 μL of appro­
time-correlated single photon counting system using pulsed light emit­ priately diluted α-amylase (10 mg/L) and 250 μL of GO suspensions
ting diodes of the EPLED as the excitation source. The wavelengths of (from 0 to 100 mg/L). The reaction was started by adding 500 μL of
excitation and emission were 280 nm and 335 nm, respectively. The gelatinized starch (1%) solution to the prepared complex solution. The
concentration of α-amylase was fixed at 300 mg/L and that of GO was solution was then incubated at 40 ◦ C for 5 min in a thermostatic water
10 mg/L. bath shaker to avoid the effect of temperature change on enzyme ac­
Synchronous fluorescence spectroscopy was applied to confirm tivity. And 1.0 mL of DNS reagent was added to stop the reaction. To
characteristic features about tyrosine (Tyr) and tryptophan (Trp) resi­ develop color, the solution was incubated in boiling water for 10 min
dues of α-amylase as a function of GO concentration. Wavelength in­ and cooled at room temperature. To remove the effect of the absorbance
tervals (Δλ) were fixed as 15 and 60 nm, which represent the Tyr and Trp of GO itself, 0.22 µm filter membranes were used to remove the GO from
residues, respectively [35]. the mixed suspensions before measuring the UV-Vis absorbance. The
The isometric three-dimensional (3D) fluorescence spectroscopy UV-Vis absorbance was measured at 540 nm using Shimadzu UV-2600
could reveal the proteins’ structural information. GO concentration was spectrophotometer. To figure out the amount of reducing sugars
set at 10 mg/L, and that of α-amylase was at 30 mg/L in all the groups. released in the reaction, a standard curve was prepared that plotted
GO and α-amylase solution were then mixed and vortexed at 200 rpm/ absorbance against amount of reducing sugars. The control set of each
min for 10 s. The excitation wavelength ranged from 200 to 500 nm, experiment was added ultrapure water instead of GO.
while the emission wavelength was in the range of 250 and 550 nm with The percent relative activities of α-amylase and α-amylase-GO
a 5 nm interval. The scan speed was 12000 nm/min, and both the complexes were calculated by the following equation [39]:
emission and excitation slits were 5 nm.
Relative activity (%) = (A1/A0)×100
2.3.4. UV-Vis absorption spectroscopy where A1 and A0 are the absorbances at 540 nm with and without GO,
Influence of GO on the UV-Vis absorption spectra of α-amylase was respectively.
recorded in the range of 200–600 nm by the UV-Vis spectrometer. GO
itself was used as the background absorbance. The α-amylase concen­ 2.3.9. Molecular docking analysis
tration was kept constant at 300 mg/L, while the GO concentration Molecular docking between GO and α-amylase was conducted to
varied between 0 and 75 mg/L. analyze their binding affinities and modes of interaction using Autodock
(http://autodock.scripps.edu/), a silico protein-ligand docking software
2.3.5. FT-IR spectroscopy [40]. The 3D coordinates of α-amylase crystal structure (PDB ID, 1C8Q;
First, 100 and 600 mg/L of α-amylase were added into the GO sus­ resolution, 2.30 Å) was obtained from the RCSB PDB database
pensions (200 mg/L) and the vials were shaken in a rotary shaking (https://www.rcsb.org/). GO structure (PubChem CID: 163320950) was
incubator at 250 rpm/min for 12 h at room temperature. The suspen­ determined from the PubChem database (https://pubchem.ncbi.nlm.
sions were then centrifuged at 10000 rpm/min for 15 min and the nih.gov) [41]. For the crystal structure of α-amylase, all water molecules
precipitates were freeze dried, which were later characterized by FT-IR. were excluded, and polar hydrogen atoms were added. After optimizing
FT-IR spectra in the 4000–400 cm− 1 wave number range were collected. the energy and adjusting the parameters of the force field, the
The amide I region (1700–1600 cm− 1) that originates from the low-energy conformations satisfying the ligand structure were selected.
stretching vibrations of C-O bond in the FT-IR spectra of the proteins was Finally, the docking results were analyzed and visualized using Dis­
analyzed by Peakfit 4.12 software. The second derivative spectra of the covery Studio Visualizer and PyMOL software.
amide I region were obtained using a 9-point Savitsky-Galey function,
which were baseline corrected and fitted to separate overlapping peaks
based on the maximum absorption intensity, band frequency, and
bandwidth from the second derivative (R2 > 0.999) [36].

3
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

3. Results and discussions analyzer, the content of O (51.99%) was the highest among the elements
of GO, fortifying the high oxidation degree of GO (Table S3).
3.1. GO characterization

As illustrated in the HR-TEM and FE-SEM images in Fig. S1, the GO 3.2. Adsorption of α-amylase on GO and interactions between them
material was consisted of wrinkled nanosheets with a thickness of single-
atom layer and a lateral size of about 100.0–2000 nm. As shown in Protein stability is critical for preserving its conformational structure
Fig. S2, the mean lateral size of GO measured by FE-SEM was 375.1 ± and biological function [49,50]. The respective UV-Vis and fluorescent
137.9 nm. The average Dh of GO in pure water was 242.0 ± 1.400 nm spectra of α-amylase in PBS did not change significantly compared with
and slightly increased to 263.0 ± 7.400 nm in the PBS solution without those in water (Figs. S3 and S4), indicating that the protein was stable
α-amylase. The zeta potential of the GO in pure water was − 48.20 ± and did not decompose under the testing conditions. The HR-TEM im­
0.3000 mV and increased to − 28.30 ± 0.8000 mV in the PBS solution, ages (Fig. S5) also reveal that GO nanosheets were well dispersed in the
implying that GO was still well dispersed in the PBS solution. Fig. 1(A) presence of 300 mg/L α-amylase. The calculated thickness of protein
illustrates that the maximum UV-Vis absorption of GO was at 227 nm, adsorption layer on GO (Table S4) according to Ohshima’s soft particle
which corresponds to the π-π * transition within GO [42]. According to theory was 3.506 nm, implying a monolayer adsorption of α-amylase on
Fig. 1(B), the value of ID/IG (the ratio of the intensity of D peak at 1343 GO surface [51].
cm− 1 to the intensity of G peak at 1576 cm− 1) in the Raman spectra was The pH did not change distinctly (~7.4) after adsorption, which
0.9431. This ratio was lower than those of preceding studies might be due to the buffering capacity of PBS. As shown in Fig. 2(A),
(1.550–1.780), demonstrating that the GO material prepared in this adsorption of α-amylase on GO surface reached saturation when
study has a relatively higher graphitization level [43,44]. The FT-IR α-amylase concentration was above 500 mg/L, and the largest adsorp­
spectrum of GO (Fig. 1(C)) reveals that there are abundant reactive tion amount was about 2000 mg/g. The adsorption isotherms are illus­
functional groups on its surface, including C– – O (1727 cm− 1), C– –C trated in Fig. 2(B). The fitted isotherms of Langmuir and Freundlich
(1624 cm− 1), C-OH (1400 cm− 1), and C-O-C (1050 cm− 1) [45,46]. The models are presented in Fig. 2(B) and S6 and the fitting parameters are
H2O molecules adsorbed on the GO surface may be responsible for the listed in Table S5. The fitting by Langmuir model (R2 = 0.9180) was
relatively wide absorption peak at 3425 cm− 1 in the FT-IR spectrum of better than that by Freundlich model (R2 = 0.8625). This suggests that
GO. The XPS spectrum of GO shows that the content ratio of C to O was the adsorption process mainly occurs via foration of monomolecular
2.02. The C 1s spectrum (Fig. 1(D)) was convolved into 4 peaks with layer on a homogeneous adsorbent surface [52], which is in agreement
binding energies of 284.4, 286.5, 287.8, and 288.8 eV, corresponding to with the calculated thickness of the protein adsorption layer. The
the C-C/C– – C, C-O-C/C-O, C– – O, and O-C– – O groups, respectively [47, calculated maximum adsorption capacity (Qe, 2317 mg/g) fitted by
48], and their contents are listed in Table S2. The oxygen-containing Langmuir model was consistent with the measured result
functional groups (i.e., C-O-C/C-O, C– – O, and O-C– – O) were predomi­ (~2000 mg/g). The Qe of α-amylase was much higher than that of other
nant with a contribution of 52.36%. As determined by the elemental proteins (e.g., 100.0 mg/g for trypsin and ~750.0 mg/g for BSA) at
similar pH and temperature, revealing that GO has stronger binding

Fig. 1. Characterization of GO: (A) UV-Vis spectrum of GO (20 mg/L in water). (B) Raman spectrum, (C) FT-IR spectrum, and (D) XPS spectrum of GO powder.

4
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

Fig. 2. (A) Adsorption of α-amylase on GO at 298 K. GO concentration was fixed at 100 mg/L and the initial α-amylase concentration varied from 50 to 1000 mg/L.
pH = 7.4. (B) Fitting adsorption isotherm of α-amylase on GO using Langmuir model. (C) Dh and (D) zeta potential of GO with and without α-amylase.

affinity with α-amylase [53,54]. Additionally, GO also exhibited much depicted in Fig. 3(B). As listed in Table S6, the correlation constants of
higher adsorption capacity to α-amylase than other nanoparticles the modified Stern-Volmer equation (Ka) decreased with increasing
(427.8 mg/g for ZnO and 276.2 mg/g for Fe3O4 nanoparticles) [55]. temperatures, suggesting that α-amylase fluorescence was statically
Hydrophobic interaction, hydrogen bonding, van der Waals force, and quenched by GO and they formed a non-fluorescent stable ground-state
electrostatic interaction force might contribute to the strong interactions complex [60].
between them due to the amphipaths of both GO and α-amylase [56]. Time-resolved fluorescence spectroscopy was further conducted to
Besides, the aromatic functional groups of both GO and α-amylase could investigate the fluorescence quenching mechanism of α-amylase caused
induce π-π stacking between them [57]. With rising concentrations of by GO. Fluorescence decay traces of α-amylase in the absence and
α-amylase, the Dh of GO increased slightly, while the GO zeta potential presence of GO are presented in Fig. 3(C) and the fluorescence lifetime of
became less negative (Fig. 2(C)(D)). Under the experimental pH (7.4), α-amylase was calculated using biexponential function [61]:
α-amylase is negatively charged, but it still contains a certain amount of
< τ >= b1 τ1 + b2 τ2
positively charged amino acid residues (e.g., lysine) [58], which may
bind to the negatively charged functional groups of GO via electrostatic where < τ > is the average lifetime of the protein; τi and bi are the fitting
attraction, lowering the absolute zeta potential of GO [59]. fluorescence lifetime and corresponding weighted proportion of fluo­
Fluorescence spectra were used to study the interaction between GO rescence quantum yield (Σbi = 1). The calculated τ values are listed in
and α-amylase, and the results are depicted in Fig. S7 and Fig. S8. The the inset table in Fig. 3(C). In dynamic fluorescence quenching, the
maximum fluorescence emission peak of α-amylase was at 340 nm, and compound elastically collides with fluorophore at its excited state and
its fluorescence spectral shape and maximum emission peak position did the lifetime of fluorophore changes. On the contrary, the fluorescence
not change after incubation with GO. GO showed negligible fluorescence lifetime of fluorophore remains unchanged if static quenching occurs
intensity, and thus the fluorescence of α-amylase-GO conjugates was and ground-state complex is formed between fluorophore and com­
primarily originated from α-amylase [18]. A decline of fluorescence pound [62]. The average fluorescence lifetime of α-amylase without and
intensity was observed as the GO concentrations rose under different with GO was similar, which directly reveals that the fluorescence of
temperatures (277 K, 298 K, and 310 K), demonstrating that the α-amylase was statically quenched by GO and they formed stable
intrinsic fluorescence of α-amylase was quenched by GO. The F/F0 of ground-state complex. This result was in good agreement with the re­
α-amylase with the addition of GO at the same concentration increased sults obtained from the modified Stern-Volmer equation [63].
with increasing temperatures (Fig. 3(A)), implying that the interaction Energy transfer between GO and α-amylase was calculated on the
was inhibited at higher temperatures. As shown in Fig. S8(D), the basis of fluorescence resonance energy transfer (FRET) theory [64]. The
variation curves of F/F0 were similar at different incubation durations, energy transfer efficiency (E), the binding distance (r) and binding dis­
suggesting that the interaction of GO and α-amylase was strong and tance taken at E = 50% (R0) were obtained by calculating the overlap
quick. integral (J) of the GO absorption spectra and the α-amylase fluorescence
The modified Stern-Volmer graphs (described in the Supplementary emission spectra (Fig. S9), and the results are shown in Table S7. The
Information) of α-amylase-GO conjugates at different temperatures are

5
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

Fig. 3. (A) F/F0 of α-amylase with increasing GO concentration at different temperatures. (B) Modified Stern-Volmer plots of α-amylase-GO system. (C) Fluorescence
decay traces of α-amylase with and without GO. τ is the fluorescent lifetime of the protein and b is the normalized pre-exponential factor. The concentrations of
α-amylase and GO were 300 mg/L and 10 mg/L, respectively. (D) Van’t Hoff plot of α-amylase-GO system.

UV-Vis spectra of GO (Fig. S9(D)) remained almost the same under The interactions of GO and α-amylase were further inspected by UV-
different temperatures (277 K, 298 K, and 310 K). If the binding dis­ Vis absorption spectra, and the adsorption spectra of α-amylase as a
tance (r) is larger than the critical distance (~10 nm), the energy function of GO concentration are displayed in Fig. S10. The obvious
transfer efficiency is too low, and the energy transfer process is hardly to peaks at 205 nm and 280 nm of α-amylase corresponded to the main
occur. The r values in this study were all below 10 nm, magnifying that chain structures of the protein and the double bonds of the Tyr and Trp
the interaction of GO with α-amylase was strong enough to allow energy residues, respectively [70,71]. The absorption intensities of α-amylase at
transfer between them and also verified fluorescence quenching 205 and 280 nm were enhanced with the increase of GO concentration,
occurred between GO and α-amylase [65]. The binding distance which verified the formation of ground-state complex between GO and
(Table S7) was positively correlated with the incubation temperatures, α-amylase and the dominance of static quenching [72].
suggesting that the energy transfer was inhibited at higher tempera­
tures. These results reveal that the α-amylase-GO complex became less
3.3. Conformational structure change of α-amylase after interaction with
stable with increasing temperatures, and energy transfer mainly derived
GO
from static quenching [66].
The Van’t Hoff equation and thermodynamic equation (described in
FT-IR spectrometry could well reveal the functional group change of
the Supplementary Information) were used to calculate the enthalpy
α-amylase before and after incubated with GO (Fig. 4(A)). After incu­
change (ΔH), free-energy change (ΔG), and entropy change (ΔS) of GO
bation of GO with α-amylase, all the peaks migrated to longer wave­
complexation with α-amylase (Fig. 3(D)), and the results are listed in
numbers and the transmittance decreased, further magnifying the
Table S8. The negative values of ΔG (− 26.71, − 26.48, and − 26.34 KJ/
formation of ground-state complex [18]. The pure α-amylase displayed
mol) and ΔH (− 29.76 KJ/mol) reveal that the interaction of α-amylase
amide A band at 3388 cm− 1, amide B band at 2929 cm− 1, amide I band
and GO was spontaneous and exothermic. The ΔH, ΔG, and ΔS
at 1653 cm− 1, and amide II band at 1536 cm− 1, representing the
(− 11.02 J mol− 1 K− 1) were all negative, indicating that hydrogen
stretching vibrations of O-H superimposed on N-H, C-H, C– – O, and C-N
bonding and van der Waals force dominated the interaction of GO and
superimposed on N-H, respectively. After incubation with GO, the amide
α-amylase [67]. The negative values of ΔG (− 26.71, − 26.48, and
B and II band at 2929 and 1536 cm− 1 in the spectrum of α-amylase
− 26.34 KJ/mol) also indicate that the binding between GO and
became smooth and even disappeared, suggesting that GO might cover
α-amylase is spontaneous and stronger than that between GO and other
the surface of α-amylase and induce rearrangement of its polypeptide
proteins (ΔG= − 0.5499 and − 0.6662 KJ/mol for HSA, ΔG=
structure [18,73,74]. When the mass ratio of GO to α-amylase increased
− 23.14 KJ/mol for human hemoglobin) [68,69].
from 1:3 to 2:1, the band at 3388 cm− 1 of α-amylase shifted obviously to

6
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

Fig. 4. (A) FT-IR spectra of GO, α-amylase, and α-amylase-GO conjugates. The concentrations of α-amylase were 50 and 300 mg/L, while the GO concentration was
fixed at 100 mg/L. (B) F/F0 in the synchronous fluorescence spectra of α-amylase (30 mg/L) as a function of GO concentration (0, 0.5, 1, 2.5, 5, 10, 25, and 50 mg/L).
The isometric three-dimensional contour spectra of (C) α-amylase and (D) α-amylase-GO conjugates. Concentrations of GO and α-amylase in the isometric three-
dimensional contour spectra were 10 and 30 mg/L, respectively.

longer wavenumber (3419 and 3420 cm− 1); while the band at intensity of Trp residue was larger than that of Tyr residues, signifying
1653 cm− 1 in the α-amylase spectrum shifted to lower wavenumber that Trp was accountable for the majority of the α-amylase intrinsic
(1637 and 1628 cm− 1). The positions of both bands shifted, implying fluorescence. The fluorescence intensity of Trp after binding with GO
that the peptide bonds of α-amylase might participate in the interaction decreased to a greater extent (from 6441 to 382.1) than that of Tyr (from
with GO. 1090 to 80.34), implying that the binding sites were closer to Trp or/and
Moreover, FT-IR as a semi-quantitative method was used to deter­ GO had stronger affinity with Trp than Tyr [76]. The emission spectra
mine the variations in the secondary structure of protein before and after peak of Tyr and Trp residues both displayed a slight red shift with the
incubated with GO, where the amide I region (1700–1600 cm− 1) cor­ increase of GO concentration, revealing that GO might cause change of
responds to the secondary structure of the protein [32]. After decon­ α-amylase conformational structure and reduce the microenvironment
volution and fitting, β-sheets (1640–1610 cm− 1), random coils hydrophobicity around Trp and Tyr residues [77].
(1650–1640 cm− 1), α-helices (1660–1650 cm− 1), and β-turns 3D fluorescence spectra can display visual conformational changes of
(~1700–1660 cm− 1) were observed [75]. After incubation of α-amylase α-amylase after incubation with GO. Fig. S13(A) reveals that GO showed
with GO, the total relative content of α-helices decreased from 31.17% little intensity of 3D fluorescence spectra. The isometric 3D contour
to 15.52%; that of β-sheets also declined from 39.15% to 29.49% spectra and projections of α-amylase with and without GO are depicted
(Fig. S11 and Table S9). In contrast, the relative abundance of random in Fig. 4 and Fig. S13, respectively. The wavelengths of the fluorescence
coils increased from 16.06% to 28.84%. These results suggest that peaks and the corresponding fluorescence intensities are listed in
α-amylase transformed into a less compact and more flexible structure, Table S10. Characteristic fluorescence peak 1 (λex/λem = 225/330) and
which might be attributed to its strong affinity with GO. peak 2 (λex/λem = 275/335) in the spectra of pure α-amylase correspond
Synchronous fluorescence spectra of α-amylase as functions of GO to Trp residue and peptide backbone structure of α-amylase, respectively
concentrations are illustrated in Fig. 4(B). The peak at 277.6 nm in the [78]. The intensity of peak 1 decreased significantly from 6586 to 2961
spectrum of α-amylase without GO represented the Trp residue after addition of GO, while the decrease of peak 2 was from 6455 to
(Δλ=60 nm) (Fig. S12(A)), whereas that at 288.8 nm corresponded to 3516. The obvious decreases of peak 1 and 2 intensities reveal signifi­
the Tyr residue (Δλ=15 nm) (Fig. S12(B)). Both peak intensities of cant changes in tertiary and secondary structures of α-amylase. Stokes
α-amylase decreased with increasing GO concentration, suggesting that shift of peak 2 remained unchanged, showing that GO did not influence
GO could bind with both amino acid residues strongly. The fluorescence the backbone of the protein distinctly. The larger Stokes shift of peak 1

7
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

after incubation with GO suggested that a red shift of the emission peak determines the biological activity of enzymes. If the secondary and
occurred and the microenvironment of Trp residues was altered by GO, tertiary structure of α-amylase is destroyed, its catalytic activity may be
which was consistent with the results of synchronous fluorescence depressed. The effect of GO on the activity of α-amylase was explored by
analysis. in vitro incubation at physiological pH of 7.4. Fig. 5(C) demonstrates
CD is an effective and sensitive method for investigating the that as the concentration of GO increased, the α-amylase activity became
conformational changes of proteins. According to Fig. 5(A), two nega­ lower. The enzymatic activity of α-amylase decreased by 9.420%,
tive characteristic peaks were observed around 208 and 222 nm in all 28.91%, and 47.02% after incubation with GO at 20, 50, and 100 mg/L,
CD spectra of α-amylase with addition of GO at different concentrations. respectively. The hydrolysis of starch catalyzed by α-amylase is related
These two peaks come from n-π * and π-π * transitions of the peptide to its conformational structure. Therefore, the inhibited enzymatic ac­
bond inside the α-helix [79], and their absolute intensities were reduced tivity might be due to the significant structural change of α-amylase
distinctly, implying the α-helix content of α-amylase decreased. The [82]. Active amino residues of α-amylase are observed on the β-strands
negative band around 217 nm representing β-sheet also exhibited a shift of the β-barrel or on loops that connect a β-strand to the surrounding
and decreased as the GO concentration increased from 0 to 75 mg/L. helix [83]. The decreased contents of α-helix and β-sheet may imply the
The secondary structure variations obtained from the CD results are destroy of α-amylase active sites and thus the loss of enzymatic activity.
consistent with the FT-IR fitting results in the amide I region. After in­ Moreover, less α-helix and β-sheet contents may also lead to looser
cubation with GO, the α-helical structure of α-amylase decreased by structure of α-amylase. As a result, GO is more likely to competitively
~11.57%, and the β-sheet content declined by ~10.02%. The results of cover the binding sites or active sites of α-amylase and substrate, pre­
FT-IR and CD both reveal that the α-helix and β-sheet contents of venting subsequent substrate from binding reaction and decreasing
α-amylase decreased due to interaction with GO. The declines of both apparent enzyme activity [84].
α-helix and β-sheet structure suggest that GO altered the secondary
structure and caused a much looser conformation of α-amylase.
ThT is a kind of cationic benzothiazole fluorescent dye with good 3.4. Molecular docking analysis
water solubility, high sensitivity, and stable affinity [80]. ThT binds
specifically with the aromatic amino acid of β-sheets to generate a The binding poses and interactions of GO (Fig. S14) with the protein
characteristic emission peak at 484 nm with the excitation wavelength are obtained with Autodock software, and the corresponding binding
of 440 nm. As displayed in Fig. 5(B), the ThT fluorescence intensity of energy was calculated. The docking pocket is a hydrophilic pocket based
α-amylase-GO conjugate at 484 nm decreased gradually as the concen­ on the interaction modes of protein and GO. Formation of hydrogen
tration of GO increased from 0 to 100 mg/L, implying that binding with bonding was observed between the amino acid residue of α-amylase (i.
GO reduced the total content of β-sheet structure [81]. e., His299, Asp300, and His305) and the oxygen or hydrogen atoms of
Conformational structure is the basis of protein function and GO, with the lowest binding energy (Fig. 6). The lowest binding energy
between α-amylase and GO was − 13.30 kcal/mol, implying that the

Fig. 5. (A) CD spectra of α-amylase (3000 mg/L) in the presence and absence of GO (0, 5, 40, and 75 mg/L). (B) ThT fluorescence of α-amylase (300 mg/L) as a
function of GO concentration (0–100 mg/L). (C) relative inhibition of enzymatic activity of α-amylase (10 mg/L) induced by GO with varying concentrations (20, 50,
and 100 mg/L).

8
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

Fig. 6. The lowest binding energy conformation of the surface binding mode of GO with α-amylase in molecular docking analysis.

binding interaction was spontaneous and a highly stable complex was Environmental implication
formed. The calculated binding energy was much lower than that be­
tween polystyrene nanoplastics and α-amylase (− 4.960 kcal/mol) and Graphene oxide (GO), as a kind of typical two-dimensional nano­
between GO and trypsin (− 10.41 kcal/mol) [18,85], revealing there is a material, is potential to accumulate in organisms and interact with
very strong binding between GO and α-amylase. Fig. 6(D) illustrates that biomacromolecules. Our study found that GO forms ground-state com­
there is a π-cation interaction between the aromatic benzene ring of GO plex with α-amylase mainly through hydrogen bonding and van der
with His305. Moreover, π-π stacking and π-π T-shaped interactions were Waals force and quenches the intrinsic fluorescence of α-amylase.
also witnessed between the aromatic benzene rings of GO with Tyr151 Consequently, the secondary conformational structure of α-amylase was
and Trp59, respectively. It was reported that Glu233, Asp197, and altered, and its enzymatic activity was inhibited. The results imply that
Asp300 residues were catalytic active sites of α-amylase [86]. The GO might greatly affect the energy metabolism and cause subsequent
interaction between Asp300 of α-amylase and GO indicates that the adverse effects to organisms.
catalytic activity of α-amylase was affected by GO. Trp59 and Asp300
are located in the α-helix secondary structure of α-amylase and their Supplementary information
binding with GO nanosheet verified the decline of α-helix content in
α-amylase after incubated with GO. In summary, both experimental and Additional methods and equations; HR-TEM and FE-SEM images of
computerized results magnify that GO binds with α-amylase strongly GO (Fig. S1); GO lateral size distribution (Fig. S2); UV-Vis spectra of
primarily via hydrogen bonding, which significantly alters the biological α-amylase in PBS and water (Fig. S3); steady-state fluorescence spectra
activity and conformational structure of α-amylase. of α-amylase in PBS and water (Fig. S4); HR-TEM images of GO with
α-amylase (Fig. S5); fitted results of Freundlich adsorption isotherm
4. Conclusions model (Fig. S6); fluorescence spectra of α-amylase as a function of GO
concentration at different temperatures (Fig. S7); fluorescence spectra of
The interactions between GO and α-amylase, as well as the confor­ α-amylase as a function of GO concentration at different incubation
mation and biological activity changes of α-amylase induced by GO, duration (Fig. S8); spectral overlap of GO absorption spectra and
were thoroughly investigated in this study. GO could interact with α-amylase emission spectra (Fig. S9); UV-Vis difference adsorption
α-amylase strongly to produce ground-state complex, and thus quench spectra (Fig. S10); curve-fitted amide I region (1700–1600 cm− 1) for
the intrinsic fluorescence of α-amylase in a static manner. The in vitro α-amylase (Fig. S11); the synchronous fluorescence spectra of α-amylase
and in silico findings reveal that GO interacts with α-amylase mainly via with and without GO (Fig. S12); the isometric 3D contour spectra of GO
hydrogen bonding, van der Waals force, and electrostatic interaction and the isometric 3D projections of α-amylase and α-amylase-GO system
spontaneously. The biological activity of α-amylase was strongly (Fig. S13); 2D structure of GO (Fig. S14); typical physiochemical prop­
inhibited by GO because of binding interaction and conformational erties of α-amylase (Table S1); results of XPS (Table S2); results of
change. This research provides novel insight into the interaction be­ elemental analysis (Table S3); calculated adsorption layer (Table S4);
tween GO and α-amylase, which reveals the potential toxicity of GO in fitting parameters of Langmuir and Freundlich models (Table S5); Ka of
physiological environment and is beneficial for risk assessment of GO in modified Stern-Volmer equation (Table S6); FRET parameters
biomedical fields. (Table S7); thermodynamic parameters of different temperatures

9
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

(Table S8); variation of relative content of secondary structures in [12] Chaudhary, K., Kumar, K., Venkatesu, P., Masram, D.T., 2021. Protein
immobilization on graphene oxide or reduced graphene oxide surface and their
α-amylase (Table S9); results of the isometric 3D fluorescence spec­
applications: influence over activity structural and thermal stability of protein. Adv
troscopy (Table S10). Colloid Interface Sci 289, 102367. https://doi.org/10.1016/j.cis.2021.102367.
[13] Tan, X., Feng, L., Zhang, J., Yang, K., Zhang, S., Liu, Z., et al., 2013.
Functionalization of graphene oxide generates a unique interface for selective
CRediT authorship contribution statement serum protein interactions. ACS Appl Mater Interfaces 5, 1370–1377. https://doi.
org/10.1021/am302706g.
Xinwei Liu: Investigation, Data curation, Writing - original draft. [14] Koeller, K.M., Wong, C.-H., 2001. Enzymes for chemical synthesis. Nature 409,
232–240. https://doi.org/10.1038/35051706.
Binbin Sun: Conceptualization, Methodology. Chunyi Xu: Visualiza­ [15] Chibata, I., Tosa, T., Sato, T., 1986. Biocatalysis: immobilized cells and enzymes.
tion, Software. Yinqing Zhang: Resources, Supervision. Lingyan Zhu: J Mol Catal 37, 1–24. https://doi.org/10.1016/0304-5102(86)85134-3.
Funding acquisition, Supervision, Writing - review & editing. Tianxu [16] Smith, S.C., Ahmed, F., Gutierrez, K.M., Frigi Rodrigues, D., 2014. A comparative
study of lysozyme adsorption with graphene graphene oxide and single-walled
Zhang: Visualization, Data curation.
carbon nanotubes: potential environmental applications. Chem Eng J 240,
147–154. https://doi.org/10.1016/j.cej.2013.11.030.
[17] Hu, W., Peng, C., Lv, M., Li, X., Zhang, Y., Chen, N., et al., 2011. Protein corona-
Declaration of Competing Interest mediated mitigation of cytotoxicity of graphene oxide. ACS Nano 5, 3693–3700.
https://doi.org/10.1021/nn200021j.
[18] Huang, S., Li, H., Liu, Y., Yang, L., Wang, D., Xiao, Q., 2020. Investigations of
The authors declare that they have no known competing financial
conformational structure and enzymatic activity of trypsin after its binding
interests or personal relationships that could have appeared to influence interaction with graphene oxide. J Hazard Mater 392, 122285. https://doi.org/
the work reported in this paper. 10.1016/j.jhazmat.2020.122285.
[19] Wu, L., Jiang, X., 2020. Enhancing peroxidase activity of cytochrome c by
modulating interfacial interaction forces with graphene oxide. Langmuir 36,
Data availability 1094–1102. https://doi.org/10.1021/acs.langmuir.9b03151.
[20] Ren, C., Hu, X., Zhou, Q., 2018. Graphene oxide quantum dots reduce oxidative
Data will be made available on request. stress and inhibit neurotoxicity in vitro and in vivo through catalase-like activity
and metabolic regulation. Adv Sci 5, 1700595. https://doi.org/10.1002/
advs.201700595.
Acknowledgments [21] Li, X., Sun, S., Guo, S., Hu, X., 2021. Identifying the phytotoxicity and defense
mechanisms associated with graphene-based nanomaterials by integrating
multiomics and regular analysis. Environ Sci Technol 55, 9938–9948. https://doi.
The authors gratefully acknowledge the National Natural Science org/10.1021/acs.est.0c08493.
Foundation of China (grants 41991313, 42207435), Ministry of Science [22] Ouyang, S., Hu, X., Zhou, Q., 2015. Envelopment-internalization synergistic effects
and metabolic mechanisms of graphene oxide on single-cell chlorella vulgaris are
and Technology (2022YFC3703200), the 111 program, Ministry of Ed­
dependent on the nanomaterial particle size. ACS Appl Mater Interfaces 7,
ucation, China (T2017002), and the China Postdoctoral Science Foun­ 18104–18112. https://doi.org/10.1021/acsami.5b05328.
dation (2020M680868). The authors would like to thank Li Meiyu from [23] Gupta, R., Gigras, P., Mohapatra, H., Goswami, V.K., Chauhan, B., 2003. Microbial
Shiyanjia Lab (www.shiyanjia.com) for time-resolved fluorescence α-amylases: a biotechnological perspective. Process Biochem 38, 1599–1616.
https://doi.org/10.1016/S0032-9592(03)00053-0.
spectroscopy test. [24] Brayer, G.D., Luo, Y., Withers, S.G., 1995. The structure of human pancreatic
alpha-amylase at 1.8 Å resolution and comparisons with related enzymes. Protein
Sci 4, 1730–1742. https://doi.org/10.1002/pro.5560040908.
Appendix A. Supporting information
[25] Whitcomb, D.C., Lowe, M.E., 2007. Human pancreatic digestive enzymes. Dig Dis
Sci 52, 1–17. https://doi.org/10.1007/s10620-006-9589-z.
Supplementary data associated with this article can be found in the [26] DeBerardinis, R.J., Thompson, C.B., 2012. Cellular metabolism and disease: what
do metabolic outliers teach us. Cell 148, 1132–1144. https://doi.org/10.1016/j.
online version at doi:10.1016/j.jhazmat.2023.131389.
cell.2012.02.032.
[27] Gupta, Jagriti, Das, Kishan, Das, Kishan, Rajamani, Paulraj, 2021. Size-responsive
References differential modulation in α-amylase by MPA-CdSe QDs: multispectroscopy and
molecular docking study. J Nanopart Res 23, 190. https://doi.org/10.1007/
s11051-021-05298-y.
[1] Wang, Y., Li, Z., Wang, J., Li, J., Lin, Y., 2011. Graphene and graphene oxide:
[28] Dhobale, S., Thite, T., Lawre, S., Rode, C., Koppikar, S., Kaul-Ghanekar, R., et al.,
biofunctionalization and applications in biotechnology. Trends Biotechnol 29,
2008. Zinc oxide nanoparticles as novel alpha-amylase inhibitors, 094907–094907
205–212. https://doi.org/10.1016/j.tibtech.2011.01.008.
J Appl Phys 104. https://doi.org/10.1063/1.3009317.
[2] Tian, B., Wang, C., Zhang, S., Feng, L., Liu, Z., 2011. Photothermally enhanced
[29] Jiang, Suisui, Li, Man, Chang, Ranran, Xiong, Liu, Sun, Qingjie, 2018. In vitro
photodynamic therapy delivered by nano-graphene oxide. ACS Nano 5,
inhibition of pancreatic α-amylase by spherical and polygonal starch nanoparticles.
7000–7009. https://doi.org/10.1021/nn201560b.
Food Funct 9, 355–363. https://doi.org/10.1039/c7fo01381g.
[3] Dhas, N., Parekh, K., Pandey, A., Kudarha, R., Mutalik, S., Mehta, T., 2019. Two
[30] Deka, J., Paul, A., Chattopadhyay, A., 2012. Modulating enzymatic activity in the
dimensional carbon based nanocomposites as multimodal therapeutic and
presence of gold nanoparticles. RSC Adv 2, 4736–4745. https://doi.org/10.1039/
diagnostic platform: a biomedical and toxicological perspective. J Control Release
C2RA20056B.
308, 130–161. https://doi.org/10.1016/j.jconrel.2019.07.016.
[31] Ernest, V., Shiny, P.J., Mukherjee, A., Chandrasekaran, N., 2012. Silver
[4] Oliveira, A.M.L., Machado, M., Silva, G.A., Bitoque, D.B., Tavares Ferreira, J.,
nanoparticles: a potential nanocatalyst for the rapid degradation of starch
Pinto, L.A., et al., 2022. Graphene oxide thin films with drug delivery function.
hydrolysis by α-amylase. Carbohydr Res 352, 60–64. https://doi.org/10.1016/j.
Nanomaterials 12, 1149. https://doi.org/10.3390/nano12071149.
carres.2012.02.009.
[5] Plachá, D., Jampilek, J., 2019. Graphenic materials for biomedical applications.
[32] Sun, B., Zhang, Y., Chen, W., Wang, K., Zhu, L., 2018. Concentration dependent
Nanomaterials 9, 1758. https://doi.org/10.3390/nano9121758.
effects of bovine serum albumin on graphene oxide colloidal stability in aquatic
[6] Kiew, S.F., Kiew, L.V., Lee, H.B., Imae, T., Chung, L.Y., 2016. Assessing
environment. Environ Sci Technol 52, 7212–7219. https://doi.org/10.1021/acs.
biocompatibility of graphene oxide-based nanocarriers: a review. J Control Release
est.7b06218.
226, 217–228. https://doi.org/10.1016/j.jconrel.2016.02.015.
[33] Sun, B., Zhang, Y., Li, R., Wang, K., Xiao, B., Yang, Y., et al., 2021. New insights
[7] Liu, J., Cui, L., Losic, D., 2013. Graphene and graphene oxide as new nanocarriers
into the colloidal stability of graphene oxide in aquatic environment: interplays of
for drug delivery applications. Acta Biomater 9, 9243–9257. https://doi.org/
photoaging and proteins. Water Res 200, 117213. https://doi.org/10.1016/j.
10.1016/j.actbio.2013.08.016.
watres.2021.117213.
[8] Yang, Y., Asiri, A.M., Tang, Z., Du, D., Lin, Y., 2013. Graphene based materials for
[34] Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of
biomedical applications. Mater Today 16, 365–373. https://doi.org/10.1016/j.
microgram quantities of protein utilizing the principle of protein-dye binding. Anal
mattod.2013.09.004.
Biochem 72, 248–254. https://doi.org/10.1006/abio.1976.9999.
[9] Zhou, Q., Ouyang, S., Ao, Z., Sun, J., Liu, G., Hu, X., 2019. Integrating biolayer
[35] Miller, J.N., 1979. Recent advances in molecular luminescence analysis. Proc Anal
interferometry atomic force microscopy and density functional theory calculation
Div Chem Soc 203–208.
studies on the affinity between humic acid fractions and graphene oxide. Environ
[36] Yang, Y., Zheng, S., Li, R., Chen, X., Wang, K., Sun, B., et al., 2021. New insights
Sci Technol 53, 3773–3781. https://doi.org/10.1021/acs.est.8b05232.
into the facilitated dissolution and sulfidation of silver nanoparticles under
[10] Zhang, P., Selck, H., Tangaa, S.R., Pang, C., Zhao, B., 2017. Bioaccumulation and
simulated sunlight irradiation in aquatic environments by extracellular polymeric
effects of sediment-associated gold- and graphene oxide nanoparticles on Tubifex
substances. Environ Sci Nano 8, 748–757. https://doi.org/10.1039/d0en01142h.
tubifex. J Environ Sci 51, 138–145. https://doi.org/10.1016/j.jes.2016.08.015.
[37] Naiki, H., Higuchi, K., Hosokawa, M., Takeda, T., 1989. Fluorometric
[11] Chen, L., Wang, C., Li, H., Qu, X., Yang, S.-T., Chang, X.-L., 2017. Bioaccumulation
determination of amyloid fibrils in vitro using the fluorescent dye, thioflavine T.
and toxicity of 13C-skeleton labeled graphene oxide in wheat. Environ Sci Technol
Anal Biochem 177, 244–249. https://doi.org/10.1016/0003-2697(89)90046-8.
51, 10146–10153. https://doi.org/10.1021/acs.est.7b00822.

10
X. Liu et al. Journal of Hazardous Materials 453 (2023) 131389

[38] Bernfeld, P., 1955. [17] Amylases, α and β. In: Methods in Enzymology. Academic [63] Ware, W.R., 1962. Oxygen quenching of fluorescence in solution: an experimental
Press, pp. 149–158. https://doi.org/10.1016/0076-6879(55)01021-5. study of the diffusion process. J Phys Chem 66, 455–458. https://doi.org/10.1021/
[39] Sun, H., Yang, B., Cui, E., Liu, R., 2014. Spectroscopic investigations on the effect j100809a020.
of N-Acetyl-l-cysteine-Capped CdTe quantum dots on catalase. Spectrochim Acta [64] Förster, Th, 1948. Zwischenmolekulare energiewanderung und fluoreszenz. Ann
Part A: Mol Biomol Spectrosc 132, 692–699. https://doi.org/10.1016/j. Phys 437, 55–75. https://doi.org/10.1002/andp.19484370105.
saa.2014.04.157. [65] Giepmans, B.N.G., Adams, S.R., Ellisman, M.H., Tsien, R.Y., 2006. The fluorescent
[40] Morris, G.M., Huey, R., Olson, A.J., 2008. Using autodock for ligand-receptor toolbox for assessing protein location and function. Science 312, 217–224. https://
docking. Chapter 8 Curr Protoc Bioinforma. https://doi.org/10.1002/ doi.org/10.1126/science.1124618.
0471250953.bi0814s24. [66] Nan, Z., Hao, C., Ye, X., Feng, Y., Sun, R., 2019. Interaction of graphene oxide with
[41] Wang, Y., Xiao, J., Suzek, T.O., Zhang, J., Wang, J., Bryant, S.H., 2009. PubChem: a bovine serum albumin: a fluorescence quenching study. Spectrochim Acta Part A
public information system for analyzing bioactivities of small molecules. Nucleic Mol Biomol Spectrosc 210, 348–354. https://doi.org/10.1016/j.saa.2018.11.028.
Acids Res 37, W623–W633. https://doi.org/10.1093/nar/gkp456. [67] Ross, P.D., Subramanian, S., 1981. Thermodynamics of protein association
[42] Loh, K.P., Bao, Q., Eda, G., Chhowalla, M., 2010. Graphene oxide as a chemically reactions: forces contributing to stability. Biochemistry 20, 3096–3102. https://
tunable platform for optical applications. Nat Chem 2, 1015–1024. https://doi. doi.org/10.1021/bi00514a017.
org/10.1038/nchem.907. [68] Khan, A., Khan, F., Shahwan, M., Khan, M.S., Husain, F.M., Rehman, Md.T., et al.,
[43] Ma, J., Liu, R., Wang, X., Liu, Q., Chen, Y., Valle, R.P., et al., 2015. Crucial role of 2021. Mechanistic insight into the binding of graphene oxide with human serum
lateral size for graphene oxide in activating macrophages and stimulating pro- albumin: multispectroscopic and molecular docking approach. Spectrochim Acta
inflammatory responses in cells and animals. ACS Nano 9, 10498–10515. https:// Part A Mol Biomol Spectrosc 256, 119750. https://doi.org/10.1016/j.
doi.org/10.1021/acsnano.5b04751. saa.2021.119750.
[44] Krishnamoorthy, K., Veerapandian, M., Yun, K., Kim, S.-J., 2013. The chemical and [69] Chakraborty, M., Mitra, I., Roy, A.J., Paul, S., Mallick, A., Das, S., et al., 2021.
structural analysis of graphene oxide with different degrees of oxidation. Carbon Contrasting spectroscopic response of human hemoglobin in presence of graphene
53, 38–49. https://doi.org/10.1016/j.carbon.2012.10.013. oxides and its reduced form: comparative approach with carbon quantum dots.
[45] Hu, X., Liu, Y., Wang, H., Chen, A., Zeng, G., Liu, S., et al., 2013. Removal of Cu(II) Spectrochim Acta Part A Mol Biomol Spectrosc 247, 119079. https://doi.org/
ions from aqueous solution using sulfonated magnetic graphene oxide composite. 10.1016/j.saa.2020.119079.
Sep Purif Technol 108, 189–195. https://doi.org/10.1016/j.seppur.2013.02.011. [70] Momeni, L., Shareghi, B., Saboury, A.A., Farhadian, S., 2016. Comparative Studies
[46] Wang, H., Liu, Y., Zeng, G., Hu, X., Hu, X., Li, T., et al., 2014. Grafting of on the interaction of Spermidine with bovine trypsin by multispectroscopic and
β-cyclodextrin to magnetic graphene oxide via ethylenediamine and application for docking methods. J Phys Chem B 120, 9632–9641. https://doi.org/10.1021/acs.
Cr(VI) removal. Carbohydr Polym 113, 166–173. https://doi.org/10.1016/j. jpcb.6b06648.
carbpol.2014.07.014. [71] Shi, J., Pan, D., Jiang, M., Liu, T.-T., Wang, Q., 2016. Binding interaction of
[47] Yang, D., Velamakanni, A., Bozoklu, G., Park, S., Stoller, M., Piner, R.D., et al., ramipril with bovine serum albumin (BSA): insights from multi-spectroscopy and
2009. Chemical analysis of graphene oxide films after heat and chemical molecular docking methods. J Photochem Photobiol B Biol 164, 103–111. https://
treatments by X-ray photoelectron and micro-raman spectroscopy. Carbon 47, doi.org/10.1016/j.jphotobiol.2016.09.025.
145–152. https://doi.org/10.1016/j.carbon.2008.09.045. [72] Toprak, M., 2016. Fluorescence study on the interaction of human serum albumin
[48] Yumitori, S., 2000. Correlation of C1s chemical state intensities with the O1s with Butein in liposomes. Spectrochim Acta Part A Mol Biomol Spectrosc 154,
intensity in the XPS analysis of anodically oxidized glass-like carbon samples. 108–113. https://doi.org/10.1016/j.saa.2015.10.023.
J Mater Sci 35, 139–146. https://doi.org/10.1023/A:1004761103919. [73] Xiao, Q., Liang, J., Luo, H., Li, H., Yang, J., Huang, S., 2020. Investigations of
[49] Yan, Z., Li, L., Li, S., Xu, Y., Yue, T., 2021. Extracellular interactions between conformational structures and activities of trypsin and pepsin affected by food
graphene nanosheets and E-cadherin. Environ Sci Nano 8, 2152–2164. https://doi. colourant allura red. J Mol Liq 319, 114359. https://doi.org/10.1016/j.
org/10.1039/D1EN00443C. molliq.2020.114359.
[50] Sun, B., Zhang, Y., Liu, X., Wang, K., Yang, Y., Zhu, L., 2022. Impacts of photoaging [74] Zhang, G., Wang, L., Pan, J., 2012. Probing the binding of the flavonoid diosmetin
on the interactions between graphene oxide and proteins: mechanisms and to human serum albumin by multispectroscopic techniques. J Agric Food Chem 60,
biological effect. Water Res 216, 118371. https://doi.org/10.1016/j. 2721–2729. https://doi.org/10.1021/jf205260g.
watres.2022.118371. [75] Xie, D., Deng, F., Shu, J., Zhu, C., Hu, X., Luo, S., et al., 2022. Impact of the frying
[51] Sun, B., Zhang, Y., Liu, Q., Yan, C., Xiao, B., Yang, J., et al., 2020. Lateral size temperature on protein structures and physico-chemical characteristics of fried
dependent colloidal stability of graphene oxide in water: impacts of protein surimi. Int J Food Sci Technol 57, 4211–4221. https://doi.org/10.1111/ijfs.15741.
properties and water chemistry. Environ Sci Nano 7, 634–644. https://doi.org/ [76] Shi, J.-H., Zhou, K.-L., Lou, Y.-Y., Pan, D.-Q., 2018. Multi-spectroscopic and
10.1039/C9EN01040H. molecular modeling approaches to elucidate the binding interaction between
[52] Latour, R.A., 2015. The langmuir isotherm: a commonly applied but misleading bovine serum albumin and darunavir, a HIV protease inhibitor. Spectrochim Acta
approach for the analysis of protein adsorption behavior. J Biomed Mater Res Part Part A Mol Biomol Spectrosc 188, 362–371. https://doi.org/10.1016/j.
A 103, 949–958. https://doi.org/10.1002/jbm.a.35235. saa.2017.07.040.
[53] Kumari, S., Sharma, P., Ghosh, D., Shandilya, M., Rawat, P., Hassan, Md.I., et al., [77] Khanna, N.C., Tokuda, M., Waisman, D.M., 1986. Conformational changes induced
2020. Time-dependent study of graphene oxide-trypsin adsorption interface and by binding of divalent cations to calregulin. J Biol Chem 261, 8883–8887. https://
visualization of nano-protein corona. Int J Biol Macromol 163, 2259–2269. doi.org/10.1016/S0021-9258(19)84464-2.
https://doi.org/10.1016/j.ijbiomac.2020.09.099. [78] Feroz, S.R., Mohamad, S.B., Bujang, N., Malek, S.N.A., Tayyab, S., 2012.
[54] Wu, C., He, Q., Zhu, A., Yang, H., Liu, Y., 2014. Probing the protein conformation Multispectroscopic and molecular modeling approach to investigate the interaction
and adsorption behaviors in nanographene oxide-protein complexes. J Nanosci of flavokawain b with human serum albumin. J Agric Food Chem 60, 5899–5908.
Nanotechnol 14, 2591–2598. https://doi.org/10.1166/jnn.2014.8521. https://doi.org/10.1021/jf301139h.
[55] Długosz, O., Matysik, J., Matyjasik, W., Banach, M., 2021. Catalytic and [79] Rigos, C.F., de, H., Santos, L., Thedei Jr, G., Ward, R.J., Ciancaglini, P., 2003.
antimicrobial properties of α-amylase immobilised on the surface of metal oxide Influence of enzyme conformational changes on catalytic activity investigated by
nanoparticles. J Clust Sci 32, 1609–1622. https://doi.org/10.1007/s10876-020- circular dichroism spectroscopy. Biochem Mol Biol Educ 31, 329–332. https://doi.
01921-5. org/10.1002/bmb.2003.494031050264.
[56] Mu, Q., Jiang, G., Chen, L., Zhou, H., Fourches, D., Tropsha, A., et al., 2014. [80] Hardy, J., Selkoe, D.J., 2002. The amyloid hypothesis of Alzheimer’s disease:
Chemical basis of interactions between engineered nanoparticles and biological progress and problems on the road to therapeutics. Science 297, 353–356. https://
systems. Chem Rev 114, 7740–7781. https://doi.org/10.1021/cr400295a. doi.org/10.1126/science.1072994.
[57] Wang, Y., Zhu, Z., Zhang, H., Chen, J., Tang, B., Cao, J., 2016. Investigation on the [81] Movaghati, S., Moosavi-Movahedi, A.A., Khodagholi, F., Digaleh, H., Kachooei, E.,
conformational structure of hemoglobin on graphene oxide. Mater Chem Phys 182, Sheibani, N., 2014. Sodium dodecyl sulphate modulates the fibrillation of human
272–279. https://doi.org/10.1016/j.matchemphys.2016.07.032. serum albumin in a dose-dependent manner and impacts the PC12 cells retraction.
[58] Habibi, A.E., Khajeh, K., Nemat-Gorgani, M., 2004. Chemical modification of lysine Colloids Surf B Biointerfaces 122, 341–349. https://doi.org/10.1016/j.
residues in bacillus licheniformis alpha-amylase: conversion of an endo- to an exo- colsurfb.2014.07.002.
type enzyme. J Biochem Mol Biol 37, 642–647. https://doi.org/10.5483/ [82] Ji, Na, Liu, Chengzhen, Li, Man, Sun, Qingjie, Xiong, Liu, 2018. Interaction of
bmbrep.2004.37.6.642. cellulose nanocrystals and amylase: its influence on enzyme activity and resistant
[59] Kubiak-Ossowska, K., Jachimska, B., Mulheran, P.A., 2016. How negatively starch content. Food Chem 245, 481–487. https://doi.org/10.1016/j.
charged proteins adsorb to negatively charged surfaces: a molecular dynamics foodchem.2017.10.130.
study of BSA adsorption on silica. J Phys Chem B 120, 10463–10468. https://doi. [83] MacGregor, E.A., 1988. α-Amylase structure and activity. J Protein Chem 7,
org/10.1021/acs.jpcb.6b07646. 399–415. https://doi.org/10.1007/BF01024888.
[60] Huang, S., Qiu, H., Lu, S., Zhu, F., Xiao, Q., 2015. Study on the molecular [84] He, T., Zhao, L., Chen, Y., Zhang, X., Hu, Z., Wang, K., 2021. Longan seed
interaction of graphene quantum dots with human serum albumin: combined polyphenols inhibit α-amylase activity and reduce postprandial glycemic response
spectroscopic and electrochemical approaches. J Hazard Mater 285, 18–26. in mice. Food Funct 12, 12338–12346. https://doi.org/10.1039/d1fo02891j.
https://doi.org/10.1016/j.jhazmat.2014.11.019. [85] Azhagesan, A., Chandrasekaran, N., Mukherjee, A., 2022. Multispectroscopy
[61] Selva Sharma, A., Ilanchelian, M., 2015. Comprehensive multispectroscopic analysis of polystyrene nanoplastic interaction with diastase α-amylase. Ecotoxicol
analysis on the interaction and corona formation of human serum albumin with Environ Saf 247, 114226. https://doi.org/10.1016/j.ecoenv.2022.114226.
gold/silver alloy nanoparticles. J Phys Chem B 119, 9461–9476. https://doi.org/ [86] Martinez-Gonzalez, A.I., Díaz-Sánchez, Á.G., de la Rosa, L.A., Bustos-Jaimes, I.,
10.1021/acs.jpcb.5b00436. Alvarez-Parrilla, E., 2019. Inhibition of α-amylase by flavonoids: structure activity
[62] Quenching of Fluorescence. In: Lakowicz, J.R. (Ed.), 2006. Principles of relationship (SAR). Spectrochim Acta Part A Mol Biomol Spectrosc 206, 437–447.
Fluorescence Spectroscopy. Springer US, Boston, MA, pp. 277–330. https://doi. https://doi.org/10.1016/j.saa.2018.08.057.
org/10.1007/978-0-387-46312-4_8.

11

You might also like