SSRN Id4270332

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

A comparative study: Green Synthesis and Evaluation of ZnO-GO and ZnO-RGO

ed
Nanocomposites for Antibacterial Applications

Reza Ahmadia, Raziyeh Fattahi Nafchib, Parvaneh Sangpour 1*b, Mozhgan Bagherib, Elahe Badieib

iew
a Department of Materials Science and Engineering, Sharif University of Technology, Tehran, Iran
b Nanotechnology and Advanced Materials Department, Materials and Energy Research Center, Iran

ev
Abstract

ZnO nanoparticle, GO and RGO sheets, ZnO-GO, and ZnO-ROG nanocomposites synthesized to

r
study their antibacterial activity. XRD and XPS confirmed that the nanocomposites successfully
synthesized. The UV-vis spectroscopy results showed the bonding between ZnO nanoparticles and GO

er
sheets can engineer the semiconductor band gap and shift the absorption edge to longer wavelengths as
compared to pure ZnO. The antibacterial activity of ZnO nanoparticles, GO, and RGO sheets after 12 h of
contact with E. coli were 84%, 81%, and 73%, respectively. Antibacterial activity for Z-1GO, Z-5GO, Z-
pe
10GO, Z-1RGO, Z-5RGO, and Z-10RGO nanocomposites were 89%, 92%, 94%, 85%, 89%, and 91%,
respectively, which are improved compared to pure ZnO (under dark conditions). Enhanced antibacterial
activity in nanocomposites may be due to the simultaneous activation of mechanisms of dissolution of Zn2+
ions from ZnO and physical interaction between the sharp edges of GO or RGO sheets and the E. coli.
ot
tn

Keywords: ZnO-GO and ZnO-RGO nanocomposite, Antibacterial activity, Antibacterial mechanisms, E.

coli
rin
ep
Pr

1Corresponding Author: sangpour@merc.ac.ir


Tel/Fax:+98 26 36280040

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
1. Introduction

ed
Nowadays, many researchers are trying to develop new antibacterial agents that are not

harmful to humans. Pathogenic bacteria grow through the formation of colonies that are resistant

iew
to traditional antibacterial methods. Therefore, designing and synthesizing new materials as

antibacterial agents is critical to solving this challenge. It has been reported that metal oxide

nanoparticles (NPs) have shown significant activity against gram-positive and gram-negative

ev
bacteria [1,2]. The metal oxides such as TiO2, ZnO, and CuO are the most critical metal oxides

with antibacterial activity [3]. ZnO NPs have been considered for many applications, including

r
solar cells, sensors, transistors, piezoelectric devices, anti-reflective coatings, and appropriate

er
antibacterial activity. ZnO NPs nanoparticles have intense antibacterial activity against a wide

range of bacteria (such as E. coli and S. aureus [4]), and ZnO is cheaper and more biocompatible
pe
than silver [5]. The photocatalytic properties of ZnO NPs are one of the antibacterial mechanisms.

In this mechanism, radicals are created by electron-holes obtained from photoexcitation. Radicals

damage cell membranes and kill bacteria by attacking organic compounds in microorganisms [6].
ot

However, the photocatalytic efficiency of ZnO NPs remains low due to the rapid recombination
tn

of the electron-hole pairs produced. Therefore, to increase the antibacterial efficiency based on

photocatalytic performance, it is necessary to delay the recombination of charge carriers. The

design and synthesis of hybrid nanomaterials based on semiconductor oxide are the most important
rin

methods for solving the challenge [1]. Researchers used graphene oxide (GO) [7] and reduced

graphene oxide (RGO) [8–10] to increase the photocatalytic properties of metal oxides.
ep

Graphene-based materials such as GO and RGO have attracted the attention of many

researchers due to their unique structure and properties [11,12]. GO and rGO have unique
Pr

applications in nanomedicine and antibacterial materials. Compared to other carbon materials such

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
as fullerene, carbon nanotubes, and graphene, these nanomaterials have less cytotoxicity [3].

ed
Semiconductor-based-GO nanocomposites have shown superior photocatalytic antibacterial

behavior under visible/UV light irradiation due to relatively lower recombination of electron-hole

iew
pairs and faster charge transfer in these photoactive [13]. Also, GO plates have sharp edges that,

through contact with bacteria, damage cell membranes and kill bacteria [14].

It has recently been reported that GO composite with CuO can increase antibacterial

ev
properties by 76% [3]. Alayande et al. [15] studied the antibacterial properties of RGO-CuO

nanocomposite films; they found that adding RGO to CuO could completely inactivate

r
Pseudomonas aeruginosa PAO1. Liu et al. [16] investigated the antibacterial activity of graphite,

er
graphite oxide, GO, and RGO. They reported that GO and RGO showed suitable antibacterial

activity against E. coli. In another study, ZnO NPs were shown to have significant antibacterial
pe
properties in contact with Staphylococcus aureus. Also, ZnO shows high biocompatibility against

human keratinocyte cells [17].


ot

This study aims to synthesize and develop novel ZnO-GO and ZnO-RGO nanocomposites

for antibacterial applications with excellent efficiency. ZnO nanoparticles interact with GO or
tn

RGO sheets through van forces in waltz, electrostatic, and hydrogen bonds. The antibacterial

properties of ZnO nanoparticles were compared with ZnO-GO and ZnO-RGO nanocomposites (1,
rin

5, and 10 %wt of GO or RGO sheets). The ZnO semiconductor nanocomposites with GO or RGO

showed improved antibacterial activity compared to pure ZnO, GO, and RGO. The highest

antibacterial activity under dark conditions after 12 h is related to Z-10GO nanocomposite, killing
ep

94% of E. coli bacteria. In nanocomposites, the bond between ZnO nanoparticles and GO or RGO

sheets leads to narrowing the energy gap and shifting the absorption edge to longer wavelengths
Pr

than pure ZnO.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
2. Materials and methods

ed
2.1. Preparation of ZnO NPs

iew
Synthesis of ZnO NPs using the green chemistry method consists of two steps. In the first

step, ionic liquid 1-butyl-3-methylimidazolium chloride was synthesized, and in the second step,

this liquid was used to synthesize ZnO NPs. For this purpose, 1-butyl chloride (Merck) and N-

methylimidazole (Merck) were poured into the balloon in a molar ratio of 1: 1.1, and the balloon

ev
was connected to the condenser. For one week, the balloon connected to the condenser was placed

under a magnetic stirrer (500 rpm) in an oil bath (60 °C). The obtained ionic liquid is in the form

r
of a concentrated phase at the bottom of the balloon. To remove unreacted raw materials

er
(immiscible with ionic liquid), the ionic liquid was cooled and washed with a diethyl ether (Merck)

solution. The diethyl ether residue was removed in an oven at 60 °C. In the second step, precursors
pe
of zinc acetate (Merck), sodium hydroxide (Merck), and synthesized ionic liquid with a molar ratio

of 1: 3: 1 were used. First, zinc acetate and sodium hydroxide were crushed separately in a mortar
ot

and then was mixed. The synthesized ionic liquid was added to the mixture of zinc acetate-sodium

hydroxide and was ultrasonicated for 30 min to obtain a homogeneous suspension. The resulting
tn

suspension was washed several times with water and ethanol and centrifuged. Finally, the

synthesized ZnO NPs were dried at 70 °C.


rin

2.2. Preparation of GO
ep

GO was synthesized by modifying the Hummer method. First, 0.5 gr of graphite powder

(Merck) and 0.3 gr of sodium nitrate (Merck) were added to 16.5 ml of cold sulfuric acid in a
Pr

balloon and then stirred (1250 rpm) for 30 min. In the next step, an ice bath was used to bring the

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
temperature to 0 °C, and 1.5 gr of potassium permanganate (Merck) was gradually added to the

ed
suspension over 30 min. Next, the 35 °C oil bath replaced the ice bath, and the suspension was

stirred for 40 min. In the next step, 70 ml of deionized water was gradually added to the system,

iew
at which point the suspension temperature reached 85 °C, and brown vapors were released from

the suspension. After completion of the reaction, 1.75 ml of hydrogen peroxide (Merck) was added

drop wise to the suspension. The resulting suspension was then washed several times with

ev
hydrochloric acid, ethanol and centrifuged (4100 rpm) for 10 min. Finally, the resulting powder

was dried under vacuum for 24 h at 60 °C.

r
2.3. Preparation of RGO er
pe
First, a suspension of GO with a 1 mg/mL concentration in deionized water was prepared

and ultrasonicated for 1 h to obtain the homogeneous suspension. Then, 1 ml of hydrazine hydrate

(Merck - 80% in water) was added to the suspension as a reducing agent. In order to complete the
ot

reduction reaction, an oil bath was used at 100 °C. At the end of the reduction reaction, the black

RGO is immersed on the water's surface and separated from the water by filtration. Finally, the
tn

synthesized RGO was washed several times with methanol and water and dried at ambient

temperature.
rin

2.4. Preparation of ZnO-GO and ZnO-RGO nanocomposites


ep

ZnO-GO and ZnO-RGO nanocomposites were synthesized by physical methods. The GO

and RGO in nanocomposites 1 %wt, 5 %wt, and 10 %wt were considered. For this purpose, three
Pr

aqueous suspensions of GO or RGO and three aqueous suspensions of ZnO were prepared. Then,

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
the suspension containing GO or RGO was added to the ZnO suspension and subjected to vigorous

ed
stirring (1400 rpm) for 3 h. After that, the suspension containing GO or RGO and ZnO was

ultrasonicated for 30 min. Finally, the suspension was centrifuged, and the ZnO-GO or ZnO-RGO

iew
nanocomposite was dried at 80 °C for 4 h. The nanocomposites were coded as Z-xGO or Z-xRGO

with x =1, 5, and 10 %wt.

2.5. Characterization methods

ev
The study of chemical bonds and functional groups in nanocomposites synthesized by

Fourier transform infrared spectroscopy (FTIR, Perkin Elmer, Spectrum RXI device in the range

r
of 400–4000 cm-1) analysis was performed. X-ray diffraction (XRD) analysis was used to evaluate

er
the crystal structure and phases of the nanocomposites (Philips PW 3710 device, Cu Kα; λ = 1.5406

lamp, voltage = 50 kV, current = 250 mA, 2θ = 5–80 and scan rate = 0.5 °/min). Scanning electron
pe
microscopy (SEM, VEGA/TESCAN model) was used to study the nanostructure of ZnO and

Oxford Stereo Scan S360 to study the nanostructure of GO and RGO. The nanoparticles' surface
ot

area, size, and porosity distribution were evaluated by measuring the adsorption isotherm of

nitrogen gas molecules using the Brunauer–Emmett–Teller (BET, Belsorp-MiNI/Gemini 2375


tn

model). The nanomaterials’ optical properties were measured using ultraviolet-visible

spectroscopy (UV-vis, PerkinElmer Spectrometer made by LambdaIs USA). The X-ray


rin

photoelectron spectrometer (XPS - Al KR anode) was used to investigate the chemical bonds

present in the nanocomposites. XPS at energy of 1486.6 eV was employed to investigate the

surface atomic composition, chemical state, and stoichiometry of the Zinc Oxide containing the
ep

GO and RGO nanosheets. Atomic force microscopy (AFM, Veeco Autoprobe CP-Research model

with 10 nm silicone tip and non-contact mode) was engaged in measuring the size and thickness
Pr

of graphene oxide sheets.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
2.6. Antibacterial activity

ed
The antibacterial activity of the synthesized samples was evaluated by the unit colony

formation (CFU) method. The LB liquid culture medium was obtained via dissolving peptone: 10

iew
g/L, salt: 10 g/L, and yeast extract: 5 g/L. After setting pH = 7, agar: 2 g is added per 100 mL of

culture medium. The LB liquid culture medium was placed in 15 mL tubes in an autoclave for 15

min at 120 °C. Then 15 mg of sample, 4.5 ml of LB liquid, and 0.5 μl of E. coli were mixed and

ev
incubated with the control sample (without an antimicrobial agent) for 12 h at 37 °C. Then diluted

samples were prepared and incubated on culture plates at 37 °C for 12 h. Finally, the number of

r
colonies formed was counted, and the percentage of antibacterial activity was calculated using

Eq.1 [3].

B‒A
er
pe
R= B
× 100 (1)

Where R: percentage of antibacterial activity, A: number of colonies counted in the sample, and

B: number of bacteria counted in the control sample.


ot
tn

3. Results and discussion

3.1. FTIR spectroscopy


rin

The FTIR spectrum for the ZnO, GO, Zn-1GO, Zn-5GO, and Zn-10GO samples are shown

in Fig. 1a. According to the FTIR spectrum of ZnO, the broad absorption peak at position 3418
ep

cm-1 belongs to the hydroxyl group in water adsorbed in the ZnO structure. The absorption peak

of the Zn-O bond is observed at 420 cm-1 wavenumbers [18]. Peaks at positions 890 cm-1, 1382
Pr

cm-1, and 1622 cm-1 are related to the bending vibrations of C-H bond, the unsymmetrical

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
stretching vibrations, and the symmetrical stretching vibrations of COO bond zinc acetate

ed
(corresponding to the remaining zinc acetate in ZnO). In general, unintense peaks are related to

zinc acetate, indicating that this substance's slight amount remains in ZnO [19]. However, the

iew
adsorption peak related to C-N binding at position 1170 cm-1 was not observed (corresponding to

ionic liquid).

According to the FTIR spectrum of GO, the unintense peak observed in position 919 cm-1

ev
is related to the C-H bending (Fig. 1a). The peaks appearing at wavenumbers 1053 cm-1, 1225

cm-1, and 1412 cm-1 correspond to the stretching vibrations of C-O bond, the stretching vibrations

r
of C-O-C bond, and the bending vibrations of O-H bond in the carboxylic acid groups. Also, the

er
peaks 1618 cm-1 and 1720 cm-1 are related to aromatic bonds C=C and carbonyl/carboxyl (C=O),

respectively. Also, the broad absorption peak related to the O-H stretching vibration is observed
pe
at position 3356 cm-1 [20]. No unknown peaks were observed in the ZnO and GO spectra,

indicating that these nanomaterials are highly pure. In addition, the FTIR spectra of Z-1GO, Z-

5GO, and Z-10GO nanocomposites are shown in Fig. 1a. In the Zn-1GO nanocomposite spectrum,
ot

the peak intensities of oxygen-containing functional groups relative to GO decreased, but with
tn

increasing GO value, the intensities of these peaks increased slightly and became similar to the

GO spectrum (for Zn-5GO and Zn-10GO nanocomposites). After bonding between ZnO and GO,

a slight shift is observed in some peaks relative to pure ZnO or pure GO, indicating that the
rin

electrostatic adsorption of ZnO on the GO plates affects the vibrational state of the GO functional

groups. For example, a displacement in the C=O peak at position 1720 cm-1 (in GO spectrum) to
ep

1650 cm-1 (in the nanocomposites spectrum) is due to the formation of a bond between zinc and

carbonyl or carboxyl groups in the GO structure [21]. Also, the nanocomposite formation causes
Pr

shifts in the Zn-O absorption band at 420 cm-1 wavenumbers to about 500 cm-1. In general, the

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
intensity reduction and shift in the nanocomposite spectrum compared to the GO and ZnO spectra

ed
indicates the successful formation of GO-ZnO nanocomposites [22].

iew
a 1622 cm-1 1382 cm-1
3418 cm-1 890 cm-1

420 cm-1
ev
1618 cm-1

r
1053 cm-1
3356 cm-1

er 1720 cm-1
919 cm-1
1225 cm-1
pe
b
ot
tn

1790 cm-1 1620 cm-1


rin

3380 cm-1
1053 cm-1
ep
Pr

Figure 1. a) FTIR spectra of ZnO, GO, Zn-1GO, Zn-5GO and Zn-10GO, and b) FTIR spectra of ZnO,

RGO, Zn-1RGO, Zn-5RGO and Zn-10RGO.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
The FTIR spectra of the ZnO, RGO, Z-1RGO, Z-5RGO, and Z-10RGO samples are shown

ed
in Fig. 1b. In the RGO spectrum, the peak intensity related to oxygen-containing functional groups

in positions 1053 cm-1, 1225 cm-1, 1412 cm-1 and 1720 cm-1 has significantly decreased compared

iew
to the GO spectrum. The reduction reaction of GO eliminates all oxygen-containing functional

groups; however, the presence of these groups is still observed in the RGO spectrum. In this study,

the formation of RGO or GO nanocomposites with ZnO is done through functional groups; thus,

ev
the functional groups must not be eliminated entirely in RGO [23], [24]. In the nanocomposite

spectrum, the peak intensities of the functional groups of RGO decreased, indicating that the

r
functional groups in RGO react with ZnO. Also, the shift at the Zn-O absorption peak in

er
nanocomposites relative to pure ZnO (from 420 cm-1 to 490 cm-1) confirms the interaction between

ZnO and RGO [25].


pe
3.2. X-Ray diffraction
ot

The XRD analysis was used for phase and structural analysis of the synthesized samples,

and the results are presented in Fig. 2. The XRD pattern obtained for pure ZnO has a Wurtzite
tn

crystal structure (hexagonal), and the impurity peak is not observed. The ZnO peaks with Wurtzite

structure are present in positions 2θ= 36.39, 40.19, 42.29, 55.72, 66.60, 74.40, and 78.84 [26],
rin

[27]. The network constants are related to the resulting structure a = b = 3.26 A° and c=5.21 A°.

The crystal size was calculated using Scherrer Eq. 2. The crystal size calculated for ZnO was about
ep

23 nm.

(0.9λ)
D = β cosθ (2)
Pr

10

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
According to the XRD pattern for GO, the peak at positions 2θ= 11.34 and 49.34

ed
corresponds to plates (002) and (100), respectively (Fig. 2). Using the Bragg equation, plate

spacing (d = 0.9 nm) was calculated for the plate (002), which increased compared to graphite (d

iew
= 0.34 nm). This could be due to oxygen-containing functional groups such as carboxyl, hydroxyl,

epoxy, and water molecules introduced between the plates [28], [29]. In the RGO pattern (Fig. 1),

the peak at positions 2θ= 24.15 and 42.59 corresponds to plates (002) and (100), respectively, in

ev
which the peak of impurity is not observed. The plate spacing (002) in RGO (d = 0.36 nm) is

relatively close to the graphite plate spacing, which is less than the GO plate spacing. This may be

r
due to removing oxygen-containing functional groups such as carboxyl, hydroxyl, and epoxy from

er
the RGO plates during the reduction process. Also, the peak shift of the plate (002) in the RGO

pattern relative to the GO confirms the change in the spacing of these plates (Fig. 2) [29].
pe
According to the XRD pattern for Z-10GO and Z-10RGO nanocomposites (Fig. 2), GO,

and RGO peaks are absent, and the pattern obtained for these nanocomposites is similar to the pure

ZnO pattern. This may be due to the low amounts of GO and RGO in the structure of these
ot

nanocomposites. Also, the network constants and crystal size in Z-10GO and Z-10RGO
tn

nanocomposites were insignificantly different from pure ZnO [30], [31].


rin
ep
Pr

11

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
ed
(101)
(100)
(002)

(110)

(103)
(102)

iew (200)
ZnO

ev
Z-10GO
(002)

r
(100)
GO
er
pe
Z-10RGO
(002)

ot
(100)
tn

RGO
rin

Figure 2. XRD pattern for ZnO, GO, RGO, Z-10GO, and Z-10RGO synthesized samples.
ep
Pr

12

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
3.3. X-ray photoelectron spectrometer

ed
Because antibacterial activity is a surface activity and depends on the surface bandings,

hence, XPS analysis was used to investigate the elemental surface chemical composition of Z-

iew
10GO nanocomposite. The XPS spectra measured for the nanocomposite are shown in Fig. 3.

According to the XPS survey of Z-10GO (Fig. 3a), the peaks of Zn (2p), C (1s), and O (1s) indicate

the presence of zinc, carbon, and oxygen elements in the nanocomposite, respectively. Therefore,

ev
there is no peak corresponding to the impurity elements in the XPS spectrum. The Zn (2p) window

is shown in Fig. 3b, in which the peak in binding energies of 1022 eV and 1045 eV corresponds to

r
Zn (2p3/2) and Zn (2p1/2), respectively. The shift of these peaks relative to the pure Zn peaks (1021

eV and 1044 eV) indicates an interaction between ZnO and GO [32], [33]. C (1s) spectra (Fig. 3c),
er
decovoluted in four peaks corresponding to carbon bondings. The peak observed at positions
pe
282.30 eV and 284.20 eV are related to C-O-Zn and C-C bonding [34], [35]. In addition, the peaks

at binding energies 285.80 eV and 288.20 eV are assigned to the C-CO and O=C-O-Zn and π-π*

bands. The presence of C-O-Zn and O=C-O-Zn peaks in the C (1s) spectrum confirms the
ot

formation of ZnO-GO nanocomposite [36]. The O (1s) spectrum is shown in Fig. 3d, that the peak

at positions 529.48 eV corresponds to oxygen ions (O2-) in the ZnO network [37]. While the
tn

presence of peaks at binding energy 531.70 eV and 532.80 eV are assigned to surface oxygen

species in ZnO-GO nanocomposite, including C-O, C-OH, and adsorbed water molecules. In
rin

addition, the presence of C-O bond indicates the binding of ZnO to GO plates. Therefore, the

results of FTIR, XRD, and XPS characterizations showed that the nanocomposite was successfully
ep

formed and surface bonding is completed for the chemical or antibacterial activity on the surface

of nanocomposite [32].
Pr

13

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
ed
a b
Zn 2p

2p3/2
Zn loss

2p1/2

iew
C Auger
O 1s

Zn 3p
C 1s

Zn loss

O 2s

ev
c d

r
Zn-O
C-O
C-C

O=C-O-Zn
C-CO
C-O-Zn er C-O
pe
ot

Figure 3. (a) XPS survey spectrum of the Z-10GO nanocomposite; high-resolution core-level XPS
tn

windows spectra of : (b) Zn (2p), (c) C (1s) and (d) O (1s).


rin

3.4. Specific surface area and porosity measurement

The high surface-to-volume ratio factor in nanomaterials makes nanomaterials more


ep

efficient than bulk materials. Therefore, measuring the specific surface area and porosity of ZnO

nanoparticles is essential for studying interaction with GO and RGO, antibacterial activity is also
Pr

a surface reaction which dependent to surface area of the nanocomposite, So, determination of

pore size and its surface area is important. BET is a suitable method for determining the specific
14

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
surface area, porosity volume, and porosity size distribution of nanomaterials with spherical,

ed
cylindrical, or layered shapes. The adsorption and desorption curves for ZnO nanoparticles as a

function of relative pressure (P/P0, P: applied pressure, and P0: saturation pressure N2) are shown

iew
in Fig. 4a. The adsorption and desorption curves in nanomaterials appear as hysteresis loops with

various shapes depending on porosity shape. The behavior of adsorption and desorption curves of

ZnO nanoparticles shows that ZnO is mesoporous nanomaterials with limited adsorption, and

ev
adsorption is stepwise (according to IUPAC standard) [38]. The data obtained from BET analysis

for ZnO nanoparticles include the average porosity diameter, porosity volume, and specific surface

r
area, which were calculated as 26.74 (nm), 0.2554 (cm3g1-), and 38.199 (m2g1-), respectively. Also,

er
the porosity size distribution of ZnO nanoparticles was obtained using the BJH theory (Fig. 4b).

According to the pore size distribution curve, ZnO nanoparticles have a wide pore size distribution
pe
(the maximum porosity size is about 46 nm). Therefore, the adsorption-desorption and the porosity

size distribution curves showed that ZnO nanoparticles are mesoporous.

a b
ot
tn
rin
ep

Figure 4. The obtained results from BET analysis for ZnO nanoparticles. a) absorption and desorption

curves, and b) porosity size distribution curves.


Pr

15

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
3.5. Microstructure study

ed
The topography of synthesized ZnO, GO, and RGO samples were evaluated using SEM,

and the resulting images are presented in Figs 5a-c. Pure ZnO nanoparticles are mostly spherical

iew
with uniform shape and particle size. The average particle size of ZnO was calculated by Image J

software to be about 60 nm (Fig. 5a). The growth mechanism of ZnO nanoparticles is divided into

the following three steps: 1) In the first step, zinc hydroxide gel is formed by zinc acetate and

ev
sodium hydroxide reaction. This step usually occurs at high molar concentrations of sodium

hydroxide. The resulting gel is complexed with 𝑂𝐻 ‒ ions, and 𝑍𝑛(𝑂𝐻)24‒ species is formed

r
according to the following reactions.

er
𝑍𝑛(𝐶𝐻3𝐶𝑂𝑂)2.2𝐻2𝑂 + 2𝑁𝑎𝑂𝐻 → 𝑍𝑛(𝑂𝐻)2 + 2𝐶𝐻3𝐶𝑂𝑂𝑁𝑎 + 2𝐻2𝑂 (3)
pe
𝑍𝑛(𝑂𝐻)2 + 2𝐻2𝑂→ 𝑍𝑛(𝑂𝐻)24‒ + 2𝐻 + (4)

2) The second step involves the formation of 1-butyl-3-methylimidazole - 𝑍𝑛(𝑂𝐻)24‒ . The 1-


ot

butyl-3-methylimidazolium chloride ionic liquid contains 1-butyl-3-methylimidazolium cation

and 𝐶𝑙 ‒ anion. Liquid ionic cations and 𝑍𝑛(𝑂𝐻)24‒ can also be absorbed by electrostatic force. 3)
tn

In the third step, the pairs of ions formed between the ionic liquid cations and 𝑍𝑛(𝑂𝐻)24‒ anion is

dehydrated, and ZnO nuclei are formed. The electrostatic attraction between the ionic liquid cation
rin

and the oxide ions in ZnO nuclei deactivated the surface of ZnO nuclei. Thus, the growth of the

nuclei stops in different directions, and the spherical structure of ZnO is observed [39], [40].
ep

According to the topography of GO sheets (Fig 5b), it is observed that the synthesized GO

has a layered structure that is formed from the accumulation of graphene sheets. Also, graphene
Pr

layers have sharp edges that can affect the antibacterial properties of GO (internal Fig. 5b). The

use of ultrasonic in the synthesis and dispersion of GO sheets reduces the number of graphene
16

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
layers (exfoliation of GO). Also, AFM image in tapping mode was used to study the microstructure

ed
of GO sheets in more detail. Tapping mode is an alternating mode of contact and non-contact mode

used to prevent damage to GO sheets. Figures 5d and e show the results obtained from AFM.

iew
According to Fig. 5e, the GO sheets have a height profile of 1.5 to 3 nm, and the width of the

sheets is about 250 nm. Comparing the height profiles of synthesized GO sheets with the

theoretical thickness of GO, observed that the GO microstructure is mainly two or three layers

ev
[41]. Also, according to the SEM and AFM images in Figs 5b and e, respectively, sharp edges in

the GO sheets are confirmed.

r
Figure 5c shows the SEM images of the RGO microstructure. There is no change in the

er
RGO microstructure compared to the GO microstructure. In the graphene oxide reduction process,

functional groups are eliminated (Fig. 1b), which reduces the distance between the graphene
pe
sheets. The RGO microstructure is made from stacking graphene sheets, that the sheets have sharp

edges (similar to the GO microstructure). Comparing the SEM images of GO and RGO, it is clear

that the size of RGO sheets is larger than that of GO sheets, which may be due to the van der Waals
ot

force between the RGO sheets, which causes graphene sheets to agglomeration in RGO [16].
tn
rin
ep
Pr

17

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
a b c

ed
es
dg
r pe
sha

iew
200 nm 500 nm 500 nm

ev
500500
nmnm 2 µm 2 µm

r
d e
er
pe
ot
tn

Figure 5. SEM images for a) ZnO, b) GO, and c) RGO. AFM results for GO, d) 2D AFM image of GO,

and e) height profile.


rin

Also, microstructural changes of nanocomposites due to the combination of ZnO with GO

and RGO were investigated using SEM. The SEM images for Z-5GO, Z-10GO, Z-5RGO, and Z-
ep

10RGO nanocomposites are presented in Fig. 6a-d (scales: 2 µm and 500 nm), respectively.

According to the SEM image of Z-5GO and Z-10GO nanocomposites (Fig. 6a and b), ZnO
Pr

nanoparticles cover the surface of GO sheets, which occurs by bonding ZnO with functional groups

18

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
on GO. ZnO nanoparticles show high accumulation on GO sheets, which by reducing the

ed
percentage of ZnO in the Z-10GO nanocomposite, ZnO accumulation is reduced (Fig. 6a and b).

According to Fig. 6c and d, ZnO nanoparticles are accumulated on GO sheets, which is due to the

iew
bonding of ZnO with a limited amount of functional groups on GO sheets. In general, the

accumulation of ZnO nanoparticles on GO sheets is higher than RGO sheets (Fig. 6). This may be

due to the presence of a more significant amount of functional groups on GO than RGO.

ev
ZnO
a nanoparticles b

r
GO sheets

er
pe
2 µm 500 nm 2 µm 500 nm

c d
ot

ZnO
nanoparticles
tn
rin

RGO sheets 2 µm 500 nm 2 µm 500 nm

Figure 6. SEM images for a) Z-5OG, b) Z-10OG, c) Z-5ROG, and d) Z-10ROG (scales: 2 µm and 500

nm).
ep
Pr

19

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
The mechanism of nanocomposite formation can be explained as follows. Through van der

ed
Waals, electrostatic, and hydrogen bond interactions, ZnO nanoparticles can be bonded to GO or

RGO sheets. The interaction between ZnO and GO or RGO can be based on hydrogen bonds

iew
between the oxygen in ZnO and the hydroxyl groups on GO. Also, ionization of functional groups

such as carboxylic acid at the edges of graphene sheets (negative charge) absorbs ZnO

nanoparticles (positive charge) through electrostatic force. In addition, the interaction between zinc

ev
atoms in the ZnO structure and oxygen atoms related to carbonyl groups in the GO or RGO

structure can affect the formation of nanocomposites [42]. Also, another mechanism involved in

r
the formation of nanocomposites is the use of ultrasonic waves in the synthesis process of

er
nanocomposites. Ultrasonic causes the formation, growth, and destruction of a large number of

microbubbles in the suspension. When these microbubbles bursts, high heat (about 5000 °K) is
pe
generated locally around the microbubbles for a few microseconds. These local heats are suitable

sites for the chemical interaction of ZnO with GO or RGO sheets [43].
ot

3.6. Ultraviolet-visible spectroscopy


tn

The study of the optical properties of the synthesized samples was performed using UV-

vis spectroscopy, and the results are presented in Fig.7. In the UV-vis spectrum of pure GO (Fig.
rin

7a), an absorption peak at wavelength 240 nm is related to the C=C aromatic bond (due to 𝜋→𝜋 ∗

transfer). Also, a shoulder at position 300 nm is observed, which is belong to the transfer of 𝑛→𝜋 ∗
ep

at C=O bonds in the carboxylic acid groups. The absorption peak and shoulder in the GO spectrum

confirmed the formation and high purity of GO [44]. In the RGO spectrum (Fig. 7b), an absorption
Pr

peak at the wavelength 267 nm corresponding to the transfer of 𝜋→𝜋 ∗ in aromatic bonds C=C.

The absorption peak in the RGO spectrum shifts to longer wavelengths than the GO spectrum,

20

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
which may be due to the removal of functional groups between the RGO sheets and the increase

ed
in conjugated bonds during the reduction process. According to the UV-vis spectrum of ZnO in

Fig. 7c, the absorption peak that appeared at position 358 nm confirmed the formation of ZnO

iew
nanoparticles [45]. In this study, the adsorption peak of ZnO nanoparticles was shifted to shorter

wavelengths than ZnO bulk (384 nm), which could be due to the smaller size of the nanoparticles.

The UV-vis spectrum of Z-1GO, Z-5GO, and Z-10GO nanocomposites are shown in Fig.

ev
7c. The bonding between ZnO nanoparticles and GO sheets leads to narrowing the semiconductor

energy gap and shifting the absorption edge to longer wavelengths than pure ZnO. This behavior

r
is due to the high electron acceptor of GO and is also observed in n-type semiconductors such as

er
TiO2. GO sheets act as electron acceptors in ZnO-GO nanocomposites, and ZnO semiconductors

act as electron donors [8]. Through the π electron rotation system, excited electrons to the
pe
conductivity level of ZnO nanoparticles are transferred to GO sheets (electron injection). The

binding of ZnO nanoparticles to GO sheets via carboxyl groups leads to shortenning the electron

transfer path from ZnO to GO [46]. Increasing GO from 1% to 10% in nanocomposites makes the
ot

energy gap narrower, and the absorption edge shifts to longer wavelengths (red-shift) (Fig. 7c).
tn

According to the UV-vis spectrum of Z-1RGO, Z-5RGO, and Z-10RGO nanocomposites in Fig.

7d, the absorption edge is shifted by increasing the percentage of RGO to higher wavelengths (red-

shift). The reason for this behavior in ZnO-RGO nanocomposites is similar to the behavior of ZnO-
rin

GO nanocomposites. Also, the red-shift in ZnO-RGO nanocomposites is higher than in ZnO-GO

nanocomposites, which may be due to the higher electron acceptor RGO than GO (Fig. 7e).
ep
Pr

21

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
ed
a b

iew
ev
c d

r
er
pe

e
ot
tn
rin
ep

Figure 7. UV–vis spectra for a) GO, b) RGO, c) ZnO, Z-1GO, Z-5GO, Z-10GO, d) ZnO, Z-1RGO, Z-

5RGO, Z-10RGO, and e) ZnO, Z-10GO, Z-10RGO.


Pr

22

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
3.7. Antibacterial activity under dark conditions

ed
The antibacterial activity of the synthesized samples was evaluated using the CFU counting

method (Eq. 1) for E. coli bacteria (for 12 h at 37 °C); the results are presented in Fig. 8. Figure

iew
8a shows a blank sample (control sample); there is no antibacterial agent in the E.coli culture

medium. Figure 8b shows the CFUs of E. coli in contact with pure ZnO nanoparticles, which kill

about 84% of the bacteria (Fig. 8k). Generally, various mechanisms have been proposed for the

ev
antibacterial activity of ZnO nanoparticles, such as the photocatalytic production of reactive

oxygen species (ROS) that destroy bacteria's cellular components or cell membranes. The energy

r
gap of the synthesized ZnO nanoparticles is in the ultraviolet range, which requires ultraviolet light

for antibacterial action (to excite electrons and create electron-hole pairs). In this study,
er
antibacterial activity was performed in a dark and ultraviolet light-free environment, suggesting
pe
that another mechanism is involved in the antibacterial properties of ZnO. The mechanism of

antibacterial activity of ZnO nanoparticles can be due to the dissolution and release of Zn2+ ions

from ZnO in the bacterial culture medium. Joe et al. [47] believed that ZnO nanoparticles attached
ot

to the bacterial cell wall dissolve, and Zn2+ ions are transfected into the cytoplasm membrane.

Lipoteichoic acid on the outer membrane of E.coli bacteria contains large amounts of negatively
tn

charged lipopolysaccharides that may provide attachment sites for ZnO nanoparticles. Therefore,

lipoteichoic can ZnO dissolution by the formation of ionic salts with Zn2+ [48]. Finally, Zn2+ ions
rin

enter the cytoplasm of E. coli through various membrane metalloproteins such as members of

ATP-binding cassette (ABC) and natural resistance-associated macrophage protein (NRAMP)


ep

[49]. The cytoplasm of E.coli disintegrates after the arrival of Zn ions; thus, the antibacterial

mechanism of ZnO nanoparticles under dark conditions is based on dissolved Zn2+ ions [5].
Pr

23

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
According to Fig. 8c and d, GO sample (81%) kills more E. coli colonies after 12 hours

ed
than RGO (71%). The antibacterial activity of GO sheets under dark conditions may be via

physical contact with the bacterium, which causes stress and destruction of the E.coli cell wall.

iew
The SEM and AFM images (Fig. 5) showed that the GO sheets had sharp edges, which destroyed

the bacterial cell membrane when bacteria came in physical contact with them. Also, Liu et al.

[16] reported that bacterial cells are severely damaged after direct contact of E. coli with graphene-

ev
based nanomaterial. Also, the oxidative stress caused by ROS has a minor role in the antibacterial

activity of GO, because graphene materials produce little superoxide. Another mechanism involves

r
the destruction of E.coli cell membrane by oxidizing the vital structure of the bacterial cell without

er
ROS intervention [50]. In this mechanism, the bacterial antioxidant (γ-L-glutamyl-L-cysteinyl-

glycine, GSH) compounds are decomposed, leading to bacterial death. GSH is a tripeptide with
pe
thiol groups and an antioxidant in bacterial cells. The thiol groups (-SH) in GSH can be oxidized

to disulfide bonds (-S-S-). The antibacterial mechanism of RGO sheets is similar to GO sheets.

The lower antibacterial activity of RGO (hydrophobic due to the removal of functional groups)
ot

than GO (hydrophilic due to functional groups) may be due to the agglomeration of RGO sheets,

which is due to the elimination of functional groups in the RGO reduction process. The van der
tn

Waals force is created between RGO sheets and prevents the graphene sheets from dispersing,

while GO sheets disperse more than RGO. Therefore, the sharp edges on RGO sheets overlap,
rin

which reduces the number of active RGO sites (reduce contact of sharp edges with bacterial

membranes) compared to GO (Fig. 8 c-d and k) [3], [51].


ep
Pr

24

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
a b c d

ed
iew
e f g

r ev
h i j
er
pe

100 k
ot

92 94
89 89 91
Antibacterial activity (%)

84 85
81
80
73
tn

60

40
rin

20

0
RGO

Z-10RGO
Z-5GO

Z-5RGO
ZnO

Z-1GO

Z-10GO

Z-1RGO
GO
ep

Figure 8. The antibacterial test images for a) blank, b) ZnO, c) GO, d) RGO e) Z-1GO, f) Z-5GO, g) Z-
Pr

10GO, h) Z-1RGO, i) Z-5RGO, j) Z-10RGO and k) antibacterial activity for all samples, after 12 h of

contact with E.coli at 37 °C.

25

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
Generally, ZnO-GO and ZnO-RGO nanocomposites show enhanced antibacterial activity

ed
compared to ZnO, GO, or RGO (Fig. 8k). According to Figs e-g, which are related to Z-1GO, Z-

5GO, and Z-10GO nanocomposites, increasing the weight percentage of GO has reduced the

iew
number of E.coli colonies. Increasing the weight percentage of GO from 1% to 10% leads to an

increase in the antibacterial activity of ZnO nanocomposites from 89 to 94% (Fig. 8k). Figures h-

j show the results of antibacterial activity of Z-1RGO, Z-5RGO, and Z-10RGO nanocomposites,

ev
which killed 85%, 89%, and 91% of E. coli bacteria after 12 h, respectively. The ZnO-GO

nanocomposites destroyed more CFUs than ZnO-RGO, which may be due to the agglomeration

r
of RGO sheets (due to the removal of functional groups). The antibacterial mechanism of ZnO-

er
GO or ZnO-RGO nanocomposites involves a combination of the following mechanisms: 1)

dissolution of Zn2+ ions from ZnO and penetration into the cytoplasm of E.coli; and 2) physical
pe
interaction between the sharp edges of GO or RGO sheets and the E. coli cell membrane.

Therefore, the increase in antibacterial activity of nanocomposites compared to pure samples is

due to the simultaneous activation of the two mechanisms mentioned in nanocomposite samples.
ot
tn

4. Conclusion

In this study, novel ZnO-GO and ZnO-RGO nanocomposites were synthesized in which
rin

ZnO nanoparticles interacted with GO or RGO sheets through van-waltz forces, electrostatic, and

hydrogen bonds. The results of XRD, FTIR, and XPS analysis confirmed that the nanocomposites
ep

were successfully synthesized. The SEM images showed that ZnO nanoparticles accumulated

more on GO than RGO due to removing oxygen-containing functional groups on RGO sheets. In

the UV-vis spectrum, increasing the GO or RGO from 1 to 10 %wt in the nanocomposites leads
Pr

to a narrowing of the energy gap, and the absorption edge shifts to longer wavelengths (red-shift)

26

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
relative to pure ZnO. The ZnO-GO and ZnO-RGO nanocomposites showed improved antibacterial

ed
activity under dark conditions compared to pure ZnO. The synthesized ZnO, GO, RGO, Zn-1GO,

Zn-5GO, Zn-10GO, Zn-1RGO, Zn-5RGO, and Zn-10RGO samples kill 84%, 81%, 73%, 89%,

iew
92%, 94%, 85%, 89%, and 91% of E. coli bacteria after 12 h at 37 °C, respectively. The enhanced

antibacterial activity of nanocomposites is due to the simultaneous activation of two antibacterial

mechanisms, including dissolution of Zn2+ ions from ZnO and physical interaction between the

ev
sharp edges of GO or RGO sheets and E. coli.

r
Declaration of interests

er
☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.
pe
☐ The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:
ot
tn

References

[1] T. Lv, L. Pan, X. Liu, T. Lu, G. Zhu, and Z. Sun, “Enhanced photocatalytic degradation of
methylene blue by ZnO-reduced graphene oxide composite synthesized via microwave-
rin

assisted reaction,” J. Alloys Compd., vol. 509, no. 41, pp. 10086–10091, Oct. 2011, doi:
10.1016/J.JALLCOM.2011.08.045.
[2] M. Kaushik et al., “Investigations on the antimicrobial activity and wound healing potential
ep

of ZnO nanoparticles,” Appl. Surf. Sci., vol. 479, pp. 1169–1177, Jun. 2019, doi:
10.1016/J.APSUSC.2019.02.189.
[3] R. Ahmadi, R. F. N. Fatahi, P. Sangpour, M. Bagheri, and T. Rahimi, “Evaluation of
Pr

antibacterial behavior of in situ grown CuO-GO nanocomposites,” Mater. Today Commun.,


vol. 28, p. 102642, Sep. 2021, doi: 10.1016/J.MTCOMM.2021.102642.

27

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
[4] H. Namazi, M. Hasani, and M. Yadollahi, “Antibacterial oxidized starch/ZnO

ed
nanocomposite hydrogel: Synthesis and evaluation of its swelling behaviours in various pHs
and salt solutions,” Int. J. Biol. Macromol., vol. 126, pp. 578–584, Apr. 2019, doi:
10.1016/J.IJBIOMAC.2018.12.242.

iew
[5] Y. Jiang, L. Zhang, D. Wen, and Y. Ding, “Role of physical and chemical interactions in
the antibacterial behavior of ZnO nanoparticles against E. coli,” Mater. Sci. Eng. C, vol. 69,
pp. 1361–1366, Dec. 2016, doi: 10.1016/J.MSEC.2016.08.044.
[6] Y. Xie, Y. He, P. L. Irwin, T. Jin, and X. Shi, “Antibacterial activity and mechanism of
action of zinc oxide nanoparticles against Campylobacter jejuni,” Appl. Environ.

ev
Microbiol., vol. 77, no. 7, pp. 2325–2331, Apr. 2011, doi: 10.1128/AEM.02149-10.
[7] H. P. Qi, H. L. Wang, D. Y. Zhao, and W. F. Jiang, “Preparation and photocatalytic activity
of Ag-modified GO-TiO2 mesocrystals under visible light irradiation,” Appl. Surf. Sci., vol.

r
480, pp. 105–114, Jun. 2019, doi: 10.1016/J.APSUSC.2019.02.194.
[8] B. Li, T. Liu, Y. Wang, and Z. Wang, “ZnO/graphene-oxide nanocomposite with

[9]
er
remarkably enhanced visible-light-driven photocatalytic performance,” J. Colloid Interface
Sci., vol. 377, no. 1, pp. 114–121, Jul. 2012, doi: 10.1016/J.JCIS.2012.03.060.
Xinjuan Liu et al., “Microwave-assisted synthesis of TiO 2 -reduced graphene oxide
pe
composites for the photocatalytic reduction of Cr( vi ),” RSC Adv., vol. 1, no. 7, pp. 1245–
1249, Nov. 2011, doi: 10.1039/C1RA00298H.
[10] M. Jabeen, M. Ishaq, W. Song, L. Xu, I. Maqsood, and Q. Deng, “UV-Assisted
Photocatalytic Synthesis of ZnO-Reduced Graphene Oxide Nanocomposites with Enhanced
ot

Photocatalytic Performance in Degradation of Methylene Blue,” ECS J. Solid State Sci.


Technol., vol. 6, no. 4, p. M36, Feb. 2017, doi: 10.1149/2.0231704JSS.
tn

[11] Y. Liu et al., “Synthesis of a poly(N-methylthionine)/reduced graphene oxide


nanocomposite for the detection of hydroquinone,” Mater. Chem. Phys., vol. 223, pp. 548–
556, Feb. 2019, doi: 10.1016/J.MATCHEMPHYS.2018.11.045.
[12] C. Chen, Z. Gan, K. Zhou, Z. Ma, Y. Liu, and Y. Gao, “Catalytic polymerization of N-
rin

methylthionine at electrochemically reduced graphene oxide electrodes,” Electrochim.


Acta, vol. 283, pp. 1649–1659, Sep. 2018, doi: 10.1016/J.ELECTACTA.2018.07.051.
[13] P. Hajipour, A. Bahrami, A. Eslami, A. Hosseini-Abari, and H. reza Hagh Ranjbar,
ep

“Chemical bath synthesis of CuO-GO-Ag nanocomposites with enhanced antibacterial


properties,” J. Alloys Compd., vol. 821, p. 153456, Apr. 2020, doi:
10.1016/J.JALLCOM.2019.153456.
Pr

[14] I. Sengupta, P. Bhattacharya, M. Talukdar, S. Neogi, S. K. Pal, and S. Chakraborty,


“Bactericidal effect of graphene oxide and reduced graphene oxide: Influence of shape of

28

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
bacteria,” Colloid Interface Sci. Commun., vol. 28, pp. 60–68, Jan. 2019, doi:

ed
10.1016/J.COLCOM.2018.12.001.
[15] A. B. Alayande, M. Obaid, and I. S. Kim, “Antimicrobial mechanism of reduced graphene
oxide-copper oxide (rGO-CuO) nanocomposite films: The case of Pseudomonas aeruginosa

iew
PAO1,” Mater. Sci. Eng. C, vol. 109, p. 110596, Apr. 2020, doi:
10.1016/J.MSEC.2019.110596.
[16] S. Liu et al., “Antibacterial Activity of Graphite, Graphite Oxide, Graphene Oxide, and
Reduced Graphene Oxide: Membrane and Oxidative Stress,” ACS Nano, vol. 5, no. 9, pp.
6971–6980, Sep. 2011, doi: 10.1021/NN202451X.

ev
[17] B. Lallo da Silva, B. L. Caetano, B. G. Chiari-Andréo, R. C. L. R. Pietro, and L. A.
Chiavacci, “Increased antibacterial activity of ZnO nanoparticles: Influence of size and
surface modification,” Colloids Surfaces B Biointerfaces, vol. 177, pp. 440–447, May 2019,

r
doi: 10.1016/J.COLSURFB.2019.02.013.
[18] S. Hosseini Largani and M. Akbarzadeh Pasha, “The effect of concentration ratio and type

er
of functional group on synthesis of CNT–ZnO hybrid nanomaterial by an in situ sol–gel
process,” Int. Nano Lett. 2016 71, vol. 7, no. 1, pp. 25–33, Dec. 2016, doi: 10.1007/S40089-
016-0197-4.
pe
[19] S. S. and, M. Ishida, and M. A. Anderson*, “Visible Luminescence and Surface Properties
of Nanosized ZnO Colloids Prepared by Hydrolyzing Zinc Acetate,” J. Phys. Chem. B, vol.
102, no. 50, pp. 10169–10175, Dec. 1998, doi: 10.1021/JP982594M.
[20] M. Veerapandian, M. H. Lee, K. Krishnamoorthy, and K. Yun, “Synthesis, characterization
ot

and electrochemical properties of functionalized graphene oxide,” Carbon N. Y., vol. 50,
no. 11, pp. 4228–4238, Sep. 2012, doi: 10.1016/J.CARBON.2012.05.004.
tn

[21] M. Nasrollahzadeh, B. Jaleh, and A. Jabbari, “Synthesis, characterization and catalytic


activity of graphene oxide/ZnO nanocomposites,” RSC Adv., vol. 4, no. 69, pp. 36713–
36720, Aug. 2014, doi: 10.1039/C4RA05833J.
[22] J. M. Lee, Y. B. Pyun, J. Yi, J. W. Choung, and W. Il Park, “ZnO Nanorod−Graphene
rin

Hybrid Architectures for Multifunctional Conductors,” J. Phys. Chem. C, vol. 113, no. 44,
pp. 19134–19138, 2009, doi: 10.1021/JP9078713.
[23] S. Elbasuney, G. S. El-Sayyad, M. Yehia, and S. K. Abdel Aal, “Facile synthesis of RGO-
ep

Fe2O3 nanocomposite: A novel catalyzing agent for composite propellants,” J. Mater. Sci.
Mater. Electron. 2020 3123, vol. 31, no. 23, pp. 20805–20815, Oct. 2020, doi:
10.1007/S10854-020-04593-Z.
Pr

[24] H. R. Azimi, M. Ghoranneviss, S. M. Elahi, and R. Yousefi, “Photovoltaic and UV detector


applications of ZnS/rGO nanocomposites synthesized by a green method,” Ceram. Int., vol.

29

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
42, no. 12, pp. 14094–14099, Sep. 2016, doi: 10.1016/J.CERAMINT.2016.06.018.

ed
[25] Q.-P. Luo, X.-Y. Yu, B.-X. Lei, H.-Y. Chen, D.-B. Kuang, and C.-Y. Su, “Reduced
Graphene Oxide-Hierarchical ZnO Hollow Sphere Composites with Enhanced Photocurrent
and Photocatalytic Activity,” J. Phys. Chem. C, vol. 116, no. 14, pp. 8111–8117, Apr. 2012,

iew
doi: 10.1021/JP2113329.
[26] J. N. Hasnidawani, H. N. Azlina, H. Norita, N. N. Bonnia, S. Ratim, and E. S. Ali,
“Synthesis of ZnO Nanostructures Using Sol-Gel Method,” Procedia Chem., vol. 19, pp.
211–216, Jan. 2016, doi: 10.1016/J.PROCHE.2016.03.095.
[27] Wali Muhammad, Naimat Ullah, Muhammad Haroon, and B. Haider Abbasi, “Optical,

ev
morphological and biological analysis of zinc oxide nanoparticles (ZnO NPs) using Papaver
somniferum L.,” RSC Adv., vol. 9, no. 51, pp. 29541–29548, Sep. 2019, doi:
10.1039/C9RA04424H.

r
[28] A. B. Alayande, C. M. Kim, J. S. Vrouwenvelder, and I. S. Kim, “Antibacterial rGO–CuO–
Ag film with contact- and release-based inactivation properties,” Environ. Res., vol. 191, p.

[29]
er
110130, Dec. 2020, doi: 10.1016/J.ENVRES.2020.110130.
X. Zhou, T. Shi, and H. Zhou, “Hydrothermal preparation of ZnO-reduced graphene oxide
hybrid with high performance in photocatalytic degradation,” Appl. Surf. Sci., vol. 258, no.
pe
17, pp. 6204–6211, Jun. 2012, doi: 10.1016/J.APSUSC.2012.02.131.
[30] M.-Q. Yang, N. Zhang, and Y.-J. Xu, “Synthesis of Fullerene–, Carbon Nanotube–, and
Graphene–TiO2 Nanocomposite Photocatalysts for Selective Oxidation: A Comparative
Study,” ACS Appl. Mater. Interfaces, vol. 5, no. 3, pp. 1156–1164, Feb. 2013, doi:
ot

10.1021/AM3029798.
[31] Y. Yang and T. Liu, “Fabrication and characterization of graphene oxide/zinc oxide
tn

nanorods hybrid,” Appl. Surf. Sci., vol. 257, no. 21, pp. 8950–8954, Aug. 2011, doi:
10.1016/J.APSUSC.2011.05.070.
[32] S. Ameen, M. S. Akhtar, M. Song, and H. S. Shin, “Vertically Aligned ZnO Nanorods on
Hot Filament Chemical Vapor Deposition Grown Graphene Oxide Thin Film Substrate:
rin

Solar Energy Conversion,” ACS Appl. Mater. Interfaces, vol. 4, no. 8, pp. 4405–4412, Aug.
2012, doi: 10.1021/AM301064J.
[33] G. Singh et al., “ZnO decorated luminescent graphene as a potential gas sensor at room
ep

temperature,” Carbon N. Y., vol. 50, no. 2, pp. 385–394, Feb. 2012, doi:
10.1016/J.CARBON.2011.08.050.
[34] H. Wang et al., “Graphene-Like Porous ZnO/Graphene Oxide Nanosheets for High-
Pr

Performance Acetone Vapor Detection,” Mol. 2019, Vol. 24, Page 522, vol. 24, no. 3, p.
522, Jan. 2019, doi: 10.3390/MOLECULES24030522.

30

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
[35] D. Fu, G. Han, Y. Chang, and J. Dong, “The synthesis and properties of ZnO–graphene

ed
nano hybrid for photodegradation of organic pollutant in water,” Mater. Chem. Phys., vol.
132, no. 2–3, pp. 673–681, Feb. 2012, doi: 10.1016/J.MATCHEMPHYS.2011.11.085.
[36] Z. Zhan, L. Zheng, Y. Pan, G. Sun, and L. Li, “Self-powered, visible-light photodetector

iew
based on thermally reduced graphene oxide–ZnO (rGO–ZnO) hybrid nanostructure,” J.
Mater. Chem., vol. 22, no. 6, pp. 2589–2595, Jan. 2012, doi: 10.1039/C1JM13920G.
[37] M. M. Hassan, W. Khan, P. Mishra, S. S. Islam, and A. H. Naqvi, “Enhancement in alcohol
vapor sensitivity of Cr doped ZnO gas sensor,” Mater. Res. Bull., vol. 93, pp. 391–400, Sep.
2017, doi: 10.1016/J.MATERRESBULL.2017.05.019.

ev
[38] Claudia Weidenthaler, “Pitfalls in the characterization of nanoporous and nanosized
materials,” Nanoscale, vol. 3, no. 3, pp. 792–810, Mar. 2011, doi: 10.1039/C0NR00561D.
[39] L. Wang, B. Zhao, L. Chang, and W. Zheng, “Fabrication of ZnO nanorods in ionic liquids

r
and their photoluminescent properties,” Sci. China Ser. B Chem. 2007 502, vol. 50, no. 2,
pp. 224–229, Apr. 2007, doi: 10.1007/S11426-007-0032-X.
[40] er
T. Alammar and A.-V. Mudring, “Sonochemical Synthesis of 0D, 1D, and 2D Zinc Oxide
Nanostructures in Ionic Liquids and Their Photocatalytic Activity,” ChemSusChem, vol. 4,
no. 12, pp. 1796–1804, Dec. 2011, doi: 10.1002/CSSC.201100263.
pe
[41] Chen Lin, Xiong-Ying Ye, and Xu-Ming Xie, “High-performance multi-functional
graphene/hexagonal boron nitride/poly(ethylene oxide) nanocomposites through enhanced
interfacial interaction by coordination,” RSC Adv., vol. 8, no. 64, pp. 36761–36768, Oct.
2018, doi: 10.1039/C8RA06325G.
ot

[42] S. Bai and X. Shen, “Graphene–inorganic nanocomposites,” RSC Adv., vol. 2, no. 1, pp.
64–98, Dec. 2011, doi: 10.1039/C1RA00260K.
tn

[43] E. K. Goharshadi, Y. Ding, M. N. Jorabchi, and P. Nancarrow, “Ultrasound-assisted green


synthesis of nanocrystalline ZnO in the ionic liquid [hmim][NTf2],” Ultrason. Sonochem.,
vol. 16, no. 1, pp. 120–123, Jan. 2009, doi: 10.1016/J.ULTSONCH.2008.05.017.
rin

[44] Hsin-Cheng Hsu et al., “Graphene oxide as a promising photocatalyst for CO 2 to methanol
conversion,” Nanoscale, vol. 5, no. 1, pp. 262–268, Dec. 2012, doi: 10.1039/C2NR31718D.
[45] J. H. Yang et al., “Growth mechanism and optical properties of ZnO nanotube by the
hydrothermal method on Si substrates,” J. Alloys Compd., vol. 475, no. 1–2, pp. 741–744,
ep

May 2009, doi: 10.1016/J.JALLCOM.2008.07.123.


[46] J. Chen, J. Shi, X. Wang, H. Cui, and M. Fu, “Recent progress in the preparation and
application of semiconductor/graphene composite photocatalysts,” Chinese J. Catal., vol.
Pr

34, no. 4, pp. 621–640, Apr. 2013, doi: 10.1016/S1872-2067(12)60530-0.

31

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332
[47] A. Joe et al., “Antibacterial mechanism of ZnO nanoparticles under dark conditions,” J. Ind.

ed
Eng. Chem., vol. 45, pp. 430–439, Jan. 2017, doi: 10.1016/J.JIEC.2016.10.013.
[48] R. Ahmadi, N. Asadpourchallou, and B. K. Kaleji, “In vitro study: Evaluation of mechanical
behavior, corrosion resistance, antibacterial properties and biocompatibility of

iew
HAp/TiO2/Ag coating on Ti6Al4V/TiO2 substrate,” Surfaces and Interfaces, vol. 24, p.
101072, Jun. 2021, doi: 10.1016/J.SURFIN.2021.101072.
[49] D. Wang, O. Hosteen, and C. A. Fierke, “ZntR-mediated transcription of zntA responds to
nanomolar intracellular free zinc,” J. Inorg. Biochem., vol. 111, pp. 173–181, Jun. 2012,
doi: 10.1016/J.JINORGBIO.2012.02.008.

ev
[50] B. Song, C. Zhang, G. Zeng, J. Gong, Y. Chang, and Y. Jiang, “Antibacterial properties and
mechanism of graphene oxide-silver nanocomposites as bactericidal agents for water
disinfection,” Arch. Biochem. Biophys., vol. 604, pp. 167–176, Aug. 2016, doi:

r
10.1016/J.ABB.2016.04.018.
[51] C. Hu et al., “GO-based antibacterial composites: Application and design strategies,” Adv.

er
Drug Deliv. Rev., vol. 178, p. 113967, Nov. 2021, doi: 10.1016/J.ADDR.2021.113967.
pe
ot
tn
rin
ep
Pr

32

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270332

You might also like