0% found this document useful (0 votes)
24 views21 pages

Fabrication of Covalently Linked Ruthenium Complex

Due to the problem of direct disposal of effluents containing antibiotics to the environment and the emergence of resistant bacterial pathogens, the wastewater treatment of the pharmaceutical industry has been known as an important research background. In this study, the refinement and photodegradation ability of one of the most widely used antibiotics, “tetracycline” was investigated by ruthenium complex immobilized on the modified graphitic carbon nitride nanotubes. For this purpose, graphitic

Uploaded by

Hossein Barani
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
24 views21 pages

Fabrication of Covalently Linked Ruthenium Complex

Due to the problem of direct disposal of effluents containing antibiotics to the environment and the emergence of resistant bacterial pathogens, the wastewater treatment of the pharmaceutical industry has been known as an important research background. In this study, the refinement and photodegradation ability of one of the most widely used antibiotics, “tetracycline” was investigated by ruthenium complex immobilized on the modified graphitic carbon nitride nanotubes. For this purpose, graphitic

Uploaded by

Hossein Barani
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 21

Fabrication of covalently linked ruthenium complex

onto carbon nitride nanotubes for the photocatalytic


degradation of tetracycline antibiotic
Mohaddeseh Shahabi Nejad 
(

[email protected]
)
https://orcid.org/0000-0002-0560-6082
Zahra Vakily 
Ali Mostafavi 
Hassan Sheibani 

Research Article

Keywords: Graphitic carbon nitride nanotubes (g-C3N4 NTs), Dichloro(p-cymene)ruthenium(II),


Photocatalyst, Tetracycline antibiotics.

Posted Date: May 16th, 2022

DOI: https://doi.org/10.21203/rs.3.rs-1016964/v2

License:


This work is licensed under a Creative Commons Attribution 4.0 International
License.
 
Read Full License

Page 1/21
Abstract
Due to the problem of direct disposal of effluents containing antibiotics to the environment and the
emergence of resistant bacterial pathogens, the wastewater treatment of the pharmaceutical industry has
been known as an important research background. In this study, the refinement and photodegradation
ability of one of the most widely used antibiotics, “tetracycline” was investigated by ruthenium complex
immobilized on the modified graphitic carbon nitride nanotubes. For this purpose, graphitic carbon nitride
nanotubes (g-C3N4 NTs) were successfully synthesized by the hydrothermal method and functionalized
with 1,10-Phenantroline-5,6-dione ligand during another step. Then, the functionalized g-C3N4 NTs were
reinforced with immobilization of dichloro(p-cymene)ruthenium(II) dimer. The structure and morphology
of the prepared photocatalyst was studied by X-ray diffraction (XRD), Fourier transform infrared (FT-IR),
scanning, and transmission electron microscopy (SEM & TEM) analyses. In the following, the
photocatalyst's ability to optically degrade the tetracycline antibiotics was performed in a suspension
reactor equipped with a LED lamp (60 W), and effective parameters such as the amount of catalyst, and
irradiation time, temperature, and pH were optimized. The results showed that the immobilization of Ru
complex onto functionalized g-C3N4 NTs improved the photocatalytic activity and increased the
degradation efficiencies to an amount of 43%. Furthermore, COD analysis has used the determination of
the amount of mineralization and results showed that the mineralization of 10 mg/L tetracycline solution
of about 90% can be performed using 20 mg of Ru (II) complex/ g-C3N4 NTs at pH = 7 after 480 min
without any additive oxidant.

1. Introduction
Tetracyclines are known as a group of broad-spectrum antibiotics that have a common basic structure
and one of the most widely used these antibiotics is tetracycline. This antibiotic is used to treat humans
and stimulate the growth of livestock and birds [1]. According to the World Health Organization (WHO),
the presence of this substance in nature is dangerous [2]. Tetracycline finds its way to surface and
groundwater through domestic and hospital effluents so livestock and poultry farms that it enters the
human food cycle by drinking water and consuming fruits and vegetables [3]. This antibiotic causes
genetic resistance and accumulates in the skeletal tissues of the human body through the formation of a
stable complex with calcium [4, 5]. Tetracycline is not biodegradable and disrupts in the biological
treatment units, and also produces more toxic products in many chemical treatment processes such as
chlorination and ozonation [6, 7]. On the other hand, the use of the surface adsorption technique only
transfers tetracycline from one phase to another and does not eliminate its contamination [8]. Therefore,
it can be said that the common removal methods of organic contaminations are inefficient and
troublesome for this antibiotic [9]. Instead, the advanced oxidation processes, especially photocatalytic
systems, are a good way to remove these compounds and are based on the production of hydroxyl
radicals [10–12]. These radicals are able to oxidize all organic compounds to the stage of carbon dioxide
and water production [13]. Most of the photocatalysts presented so far are efficient only in the ultraviolet
range [14, 15], and since only a small percentage of sunlight is made of ultraviolet light and most of that

Page 2/21
is visible light, so it needs to design and fabrication of the new generation of active photocatalysts in the
visible range. In recent years, a lot of research has been done on semiconductors that have a suitable gap
and can be used in the visible range [16–18]. Graphite nitride carbon is one of the semiconductors that
has a smaller bandgap (∼2.7 eV) than many other common photocatalysts such as titanium oxide and
zinc oxide, and as a result, it is more efficient at absorbing sunlight [19, 20]. This substance is non-toxic,
cheap and easy to prepare [21]. However, the most important weakness of g-C3N4, which reduces its
performance in optical degradation, is the rapid recombination of electron-hole pairs [22, 23]. Increasing
efficiency in the optical degradation operations depends on the degree of separation of electron-hole
pairs. To produce active oxygen species, it is necessary that the electrons and produced holes can react
separately with the water or oxygen molecules around them [24]. There are two major ways to solve this
g-C3N4 problem and achieve maximum performance; First, changes in the morphology and particle size
of the photocatalyst: The particle size of a photocatalyst has a large effect on the energy gap. As the
particle size decreases along with increasing the width of the energy gap, and so its oxidation-reduction
strength will increase [25]. Second, the decoration of g-C3N4 with other conductive and semiconductor
metals; these metals increase the absorption of visible light and the efficiency of the photocatalyst by
producing hot electrons and injecting them into the semiconductor [26].

In the present project, the above two techniques were used to fabricate an efficient photocatalyst in the
visible range. In this regard, the g-C3N4 nanotubes were synthesized instead of g-C3N4 bulk and it cause
to improve the energy band gap and reduce the recombination rate of the electron-hole pair. In addition,
the use of g-C3N4 nanotubes increases the surface area and has a positive effect on photocatalyst
performance. To implement the second approach and further enhance the photocatalytic properties, the
surface of the nanotubes was decorated with a ruthenium complex. The ruthenium complexes have the
ability to absorb light in the wavelength range of 400–800 nm and cause the formation and injection of
electrons into the photocatalytic system. In the end, the photocatalytic ability of Ru(II)complex/g-C3N4
NTs was investigated in the decomposition of tetracycline antibiotic under visible light and the effect of
factors such as radiation time, amount of photocatalyst, pH, and temperature on degradation efficiency
was investigated. The advantages of this method are maintaining the photocatalyst stability and
reusability in the heterogeneous system.

2. Experimental
2.1. Material and methods
Cyanoric cloride (99%), and 1,10-Phenanthroline (99.5%) were purchased from Merck- Millipore chemical
Co. and used as received. Melamine (99%), Dichloro(p-cymene)ruthenium(II) dimer (99%), and
Tetracycline (98%) were obtained from Sigma-Aldrich. The prepared Ru(II)complex/g-C3N4 NTs was
characterized by fourier transform infrared (FT-IR, Bruker, Germany, RT-DLATGS detector) spectroscopy,
scanning electron microscopy (FE-SEM, TESCAN-MIRA3), transmission electron microscope (TEM,
EM10c-100 KV) and the concentration of tetracycline solutions have been monitored using Ultraviolet-

Page 3/21
visible (UV-Vis) spectroscopy was performed in the range of 200–900 nm wavelength on a Cary 50 single
detector double beam in time spectrophotometer (Varian, Australia).

2.2. Synthesis of g-C3N4 NTs


The preparation of g-C3N4 NTs was done in two separately steps. First, white crystals of melamine-
cyanuric acid were prepared by a hydrothermal method briefly, 0.5 mmol of melamine and 0.5 mmol of
cyanuric chloride were dissolved in 75 mL deionized water and kept stirring for 30 min. The resultant
transport solution was transferred to an autoclave flask and maintained at 180°C for 6 h. After this time,
the autoclave was cooled until to ambient temperature and the obtained white needle-like crystals were
centrifuged and washed several times with deionized water to remove organic impurities and freezer
dried for 24 hours.

The second step was the calcination process as follows the melamine-cyanuric chloride crystals were
calcined at 450°C for 5 h with a heating rate of 1°C min− 1 under nitrogen inert atmosphere.
2.3. Functionalization of g-C3N4 NTs with phenantroline-5,6-
dione
In order to functionalize graphitic carbon nitride nanotubes, 250 mg of g-C3N4 NTs along with 1 mmol of
phenantroline-5,6-dione ligand were dispersed in 50 ml of acetonitrile and after equipping the reaction
vessel with the condenser it was refluxed for 60 hours at 60 ° C. At the end of the reaction time, the
functionalized g-C3N4 NTs were collected by centrifugation and dried at 50 ° C after washing several
times with acetonitrile.

2.4. Synthesis of Ru (II) complex/ g-C3N4 NTs


A 250mL flask was charged with the g-C3N4 NTS bearing the phenantroline-5,6-dione ligand (250 mg),
[RuCl2(p-cymene)]2 (0.2 mmol) and 200 mL anhydrous dichloromethane. The obtained suspension was
refluxed for 2 h at 50°C. After that, the resultant product was separated through decanting of the solvent
and the nanotubes were washed several times with dried dichloromethane to remove extra values of
ruthenium moieties. according to ICP-OES analysis amount of Ru was 10% in the obtained photocatalyst.

2.5. Study of the photocatalytic activity of Ru (II) complex/


g-C3N4 NTs
All photocatalytic reactions were performed in a 50 ml beaker containing 25 ml of tetracycline solution
(10 mg/L). During irradiation of visible photons, the sample containing the nanocomposite was stirred by
a magnetic stirrer to homogenize the system and prevent photocatalytic precipitation. The 60-watt visible
light source was placed directly on top of the beaker at a distance of 15 cm from and at a distance of 15
cm from it, and despite the heat generated by the lamp, no cover was used. At the beginning of the test,
from the moment that the samples were immersed in the solution, placed in the dark for 30 minutes to

Page 4/21
reach the adsorption /desorption equilibrium. The concentration of initial and residual tetracycline in the
samples that irradiated by UV-Vis spectrometer at different times were measured after separating of
nanocomposite by centrifuge and then the percentage of drug removal was calculated according to Eq. 1.

Photocatalytic destruction (%) = (C0-C)/C0 × 100 (1)

In this equation, C0 is the initial concentration of tetracycline and C is the concentration of the tetracycline
remaining in the solution, after the photodegradation reaction. It should be noted that all experiments
were repeated twice and the average value was reported. In addition, the COD of the samples was
measured every hour by FAS titration with potassium permanganate [27].

2.6. Adsorption kinetics study


The kinetic of photocatalytic degradation of pesticides in the heterogeneous oxidation systems with
visible light follows the Langmuir-Hinslowwood kinetic model (Eq. 2) [28–30].

dC kKC
r= = (2)
dt 1 + KC

Where r, C, t, k and K are the oxidation rate (mg/L min), herbicide concentration (mg/L), irradiation time
(min), reaction rate constant (1/min) and reaction adsorption coefficient (L/mg), respectively. At low
initial concentrations of herbicides, Eq. 2 is changed to the quasi-first-order equation (Eq. 3).

ln
( )
C
C0
= kKt = k appt(3)

By plotting ln (C/C0) versus t, the slope of the diagram shows the apparent rate of photocatalytic
degradation.

3. Results And Discussion


3.1. Investigation of Synthesis of Ru (II) complex/ g-C3N4
NTs
So far, various methods have been introduced for the synthesis of g-C3N4, all of which use nitrogen-rich
precursors including cyanamide, dicyandiamide, melamine, urea, thiourea, triazine derivatives, and
heptazine [31, 32]. One of the simplest and most common methods for synthesizing this substance is the
condensation method [33–36]. In this method, the mentioned precursors are converted to g-C3N4 at 550 °
C. But the main challenge of this method is the variable condensation efficiency, which leads to the
production of a product with a small surface area and, of course, limits its application to photocatalytic
processes [37]. However, the surface area of g-C3N4 depends on the type of precursor and synthesis

Page 5/21
conditions, and by changing the above factors, a high surface area can be achieved. Typically, the use of
melamine precursors results in the production of g-C3N4 with a high N/C ratio [38]. The higher the N/C
ratio, the lower the energy bandgap and the higher the photocatalytic property. If melamine is used in
combination with an oxygen-containing precursor to making g-C3N4, the product has more porosity and a
higher surface area [39]. The use of chlorine-containing precursors has a similar effect, increasing the
porosity and surface area, possibly due to the evaporation of chlorine ions in the form of HCl [40]. The
second approach to achieve a high surface area is to change the synthesis conditions to create
nanostructures. Most reports on nanostructures derived from graphite carbon-nitride are about
nanoporous and nanoplate [41, 42]. In addition to these two structures, nanotubes have recently been
introduced as a new morphology of g-C3N4, which, due to their high specific surface area, regular
crystalline walls available for mass transfer, have desirable semi-conductivity properties [43]. There are
two general methods for synthesizing g-C3N4 NTs: First method, hard templating method [44], which is
done by polymerizing the precursors in the created space of the template (such as silica) and then
removing the template by reagents that dissolve the silica. Although the g-C3N4 NTs obtained from this
method has a high surface area, its crystallinity has decreased. Crystallinity is an effective parameter
against the separation of excited charges in the photocatalytic process [45]. The second method is the
soft templating [46], which its general mechanism is based on the polymerization of nitrogen-rich
precursors in the space that has been created by soft template such as trithion X10, P12, and F124 [47].
The disadvantage of this method is the high polymerization temperature, which leads to the loss of the
template before the complete formation of the nanotube. In the present project, g-C3N4 NTs were
synthesized by the joint polymerization of nitrogenous (melamine) and chlorinated (cyanuric chloride)
precursors through the hydrothermal process without the use of any template. The cyanuric chloride is
unstable under aqueous conditions therefore; it was converted to cyanuric acid as an oxygenated
precursor created melamine-cyanurate crystals through a series of hydrogen bonds with melamine.
Finally, the growth of these crystals within 6 hours and calcination at 450°C, it cause to produce g-C3N4
NTs. Because of incomplete polymerization of melamine, there are free amine groups at the outer edges
of the prepared g-C3N4 NTs and these defects in the structure cause to balance in the energy bandgap
and also it facilitates the functionalization of the surface. So here, the functionalization was done in the
through formation imine band between carbonyls of phenantroline-5,6-dione ligand and free amines of g-
C3N4 NTs. In the end, the coordination of immobilized bidentate ligands with dichloro(p-
cymene)ruthenium(II) led to the stabilization of the relevant complex on the outer edges of g-C3N4NTs.
Scheme1 depicted the steps of the synthesis of the photocatalyst.

3.2. Catalyst characterization


3.2.1. Fourier transform-infrared spectroscopy (FT-IR)
The functional groups of melamine-cyanuric acid, g-C3N4 NTs, Ru(II)complex/g-C3N4 NTs samples were
confirmed by FTIR spectroscopy. The melamine-cyanuric acid spectrum (Fig. 1a) is extremely crowded,

Page 6/21
which can be attributed to the formation of a complicated network of hydrogen bonds. The appeared
peaks at 3088, 3230, and 3391cm− 1 are due to the symmetric stretching vibrations of the amine groups
of melamine cations that have shifted to shorter wavelengths (blue shift) than pure melamine [48–49].
The asymmetric stretching vibrations of the amine groups attributed to bored peaks around 2500–2700
cm− 1. In additional, the intense peak at 1731 cm− 1 is assigned to carbonyl stretching vibrations of
cyanurate anions which produced from hydrolysis of cyanuric chloride. The strong peaks in region 1400–
1700 cm− 1 corresponded to bending vibrations of N-H and stretching vibrations of C-N and C = N. These
peaks in the assembly indicate the successful synthesis of melamine-cyanuric acid crystals. After
pyrolysis treatment, the arrangement of the absorption bands was changed (Fig. 1b). For example, the
intensity of absorption bands attributed to N-H groups dropped sharply in region > 3000 cm− 1, and also
the assigned peak of carbonyl at 1731 cm− 1 disappeared which confirmed the release of NH2 and H2O
during pyrolysis treatment to the formation of g-C3N4. Figure 1c again shows the decreasing of the
vibrations of the amino groups, which, together with the new appeared band at 1650 cm− 1, proves the
chemical modification of the nanotubes through the formation of the imine band with phenantroline-5,6-
dione ligand.

3.2.2. Ultraviolet and visible absorption spectroscopy (UV-


Vis)
To get the UV-Vis absorption spectrum of g-C3N4 Nts, this material was dispersed in ethanol by ultrasonic
bath for 20 minutes and then its adsorption was measured. As shown in Fig. 2, g-C3N4 NTs have
maximum adsorption at 368 nm, which has a blue shift (the transition to a higher energy level or shorter
wavelength) relative to the bulk sample [50]. In nanomaterials, the distance between the electron-hole is
controlled by the particle size, so that as the particle size shrinks to the nanometer scale, the movement
of excitons are limited and a transition in the optical spectrum is observed [51]. In general, the transfer of
optical spectrum to higher energies due to the decreasing of particle size means an increase in the
forbidden band energy. The energy difference (in units of electron volts) between the highest valence
band and the lowest conduction band is called the energy bandgap [52]. To calculate the bandgap energy
of a material, the absorption coefficient parameter of the material must first be obtained. The absorber
coefficient is an important parameter in optical applications and is calculated for a transparent thin film
from the following relation:

1 1
 α = d ln T   ( 4)

Here, d is the thickness of the thin film in nanometers and T is the percentage of light transmission. Since
g-C3N4 has a direct energy gap, so by direct electron transfer between the capacitance band and the
conduction band, the band gap energy (Eg) is obtained from the Tuac Eq. (5).

(𝛼ℎ𝜈)(1/ 𝑛) = 𝐴(ℎ𝜈 − 𝐸𝑔) (5)

Page 7/21
Here, A is constant, α is the absorption coefficient, Eg is the width of the energy gap, and hν is the energy
of the emitting photon with respect to the absorption spectrum of the synthesized g-C3N4 NTs. Also, n for
g-C3N4 NTs with a straight and permissible bandgap is equal to 1/2. According to Fig. 2, the energy gap
of g-C3N4 NTs can be calculated by drawing (αhν)2(eV.cm− 1)2 in versus of hν (eV) and extrapolation its
linear part. This energy gap was obtained 2.96 eV.

3.2.3. X-ray diffraction (XRD)


The prepared melamine-cyanuric acid sample and its Ru(II)complex/g-C3N4 NTs have been characterized
by X-ray powder diffraction technique (Fig. 3). The presence of sharp diffraction peaks in the XRD of the
synthesized melamine-cyanuric acid sample during the hydrothermal process confirms that the obtained
product is highly crystalline [53]. The X-ray diffraction pattern changed completely after the pyrolysis
process and the formation of g-C3N4 NTs nanotubes, followed by chemical modification.

New peaks have appeared around 2θ = 13.1 and 27.2 and confirm the formation of graphite nitride
carbon phases in the sample [54]. An important point that can be seen in this image is the decrease in the
intensity of the peaks after heat treatment, which indicates the percentage of crystallinity of the sample is
reduced, or in other words, the long-range order of the structure is reduced.

During the heat treatment, the carbon nitride surfaces appear to be oxidized and the layers separated.
Also, due to the oxidation process, the structure is defective and its effects in reducing the peak intensity
of this material have appeared. Another reason for the low intensity of peaks is the modification process
when an element enters the structure of the compound through chemical bonding, causing the
corresponding peaks to flatten or shift and due to the stabilization conditions of the complex, there is no
possibility of ruthenium phases in the final structure.

3.2.4. Scanning Electron Microscopy (SEM)


Scanning electron microscopy provides useful information from topography and morphology of the
prepared photocatalyst. Scanning electron microscope images of the sample are shown in Fig. 5. These
images were taken with different magnifications to determine the overall morphology of the sample and
how the nanotubes are arranged. As you can see, the study sample contains clusters of orderly
nanotubes and the amount of unwanted particles in that is very low.

To determine the composition of the elements in the prepared photocatalyst, the distribution of active
sites, and to determine whether the ruthenium complex is embedded in the surface of the nanotube
matrix or in the inner part of the matrix, two complementary analyzes of energy-dispersive X‐ray pattern
and element distribution map (EDS & Map) were used. The EDS spectrum confirmed the presence of the
elements carbon, nitrogen, ruthenium, and chlorine with the appearance of the corresponding peaks and
the mapping analysis of polymer shows uniform distribution and non-accumulation of elements that
enhance the properties of the photocatalyst.

Page 8/21
3.2.5. Transmission electron microscopy (TEM)
Transmission electron microscopy (TEM) images showed that the structure of g-C3N4 is in the form of
hollow curved tubes with an outer diameter of less than 10 nm and a length of several hundred
nanometers (Fig. 5). It appears that after chemical functionalization of nanotubes and loading of
ruthenium (II) complex, failure has occurred in areas where defects such as pentagons and heptagons. In
addition, the modification process has resulted further distance between the nanotubes, which has a
constructive effect on increasing the surface area of the photocatalyst. The Ru(II) complex/g-C3N4 NTs
show accumulations in some areas, which possibly due to the agglomeration of graphitic nanotubes.
This re-stacking of graphitic nanotubes is dependent on strong van der Waals forces and π-π interactions
of sp2 structures and hydrogen bonding.

3.3. Investigation of effective factors in optical degradation


of tetracycline
3.3.1. Influence of amount of photocatalyst
As shown in Fig. 6, the degradation efficiency of tetracycline increased with increasing photocatalyst
concentration to 30 mg, after which the degradation efficiency remained almost constant. This can be
attributed to the fact that when all the antibiotic molecules landed on the nanotubes, the excess amount
of photocatalyst due to the lack of antibiotic molecules had no effect on the percentage of degradation
and even slightly reduced the reaction rate. The decreasing in reaction rate at a dose of 40 mg was
caused by two main factors. First; g-C3N4 nanotubes tend to agglomeration due to their nanometric size
and high surface energy, and when the concentration of these nanotubes exceeds a certain limit, the
activated nanotubes will be deactivated by contact with the base catalyst, and to follow it, the catalytic
efficiency decreases. Second, with the increasing amount of photocatalyst, the turbidity increased and the
light scattering occurs due to the collision of optical rays with the catalyst particles scattered in the
solution and a number of light photons lose their energy and thus the efficiency of photocatalytic
processes decreases.

3.3.2. Influence of irradiation time


To investigate the effect of irradiation time on degradation efficiency, a certain concentration of
tetracycline was exposed to an optimized amount of photocatalyst (30 mg) and a frequency of 60 watts
at different times. The results are shown in Fig. 7. As you can see, the percentage of antibiotic
degradation increased with increasing reaction time, so that in 90 minutes, more than 99% of the
tetracycline molecule was broken down into smaller components. This step was followed by monitoring
the UV spectrum. But the degradation of the small organic components until producing carbon dioxide
and water takes longer so that after 6 hours, 30% of the smaller molecules are still present in the solution.
Chemical Oxygen Demand (COD) technique was used to determine the concentration of organic
components in the second step. In this method, the sample is strongly refluxed by acidic solutions and a
Page 9/21
certain amount of potassium dichromate (K2Cr2O7). After digestion of the sample to determine COD, the
residual and unoxidized amount of potassium dichromate is titrated with ammonium sulfate to
determine the amount of potassium dichromate consumed and the oxidized material is calculated in the
oxygen equations.

3.3.3. Influence of temperature


The influence of temperature on tetracycline degradation was shown in Fig. 8. Changes in process
velocities in the range of 0 to 75 ° C indicate that degradation rate is directly related to reaction
temperature (Fig. 8). In fact, the reaction temperature, on the one hand, contributes to the production of
dissolved O2 and the degradation of H2O2, which ultimately leads to the production of active hydroxyl
radicals, and on the other hand, it provides the activation energy for the reaction. Based on this, it can be
said that the catalytic process is exothermic and by drawing ln k (rate constant) versus the inverse of the
temperature (kelvin) and in accordance with Arrhenius equation [55] (Eq. 6), the activation energy of the
reaction was obtained from the slope of the straight line. The value of the slope is equal to -Ea/R where R
is a constant equal to 8.314 J/mol K. (Fig. 9).

Ea
lnk = lnA − (6)
RT

The activation energy (Ea) of tetracycline degradation is estimated to be 0.94 kJ/mol.

3.3.4. Influence of pH
In optical degradation processes, pH can affect the desired decomposition rate of the contaminant.
Previous studies have shown that pH plays an important role in the breakdown and elimination of
antibiotics. As shown in Fig. 5, there is a significant difference between the percentage of tetracycline
degradation at different pHs and the rate of decomposition at alkaline pH is significantly higher and
optimized at pH = 9. The pH variable affects the adsorption and dissociation capacity of the target
compounds, the electric charge distribution on the surface of the catalysts, and the oxidation potential of
the conduction band.

Since the isoelectric point of g-C3N4 based materials is around pH = 9 [56] and the tetracycline antibiotic
acidic, the effect of pH on the photocatalytic process can be justified by the presence of electrostatic
forces between the surface of g-C3N4 NTs and tetracycline. Thus, at pHs > 9, the g-C3N4 NTs surface has a
negative charge and the tetracycline molecules also have a negative charge, therefore, the force between
them is repulsive and thay don’t have any reluctant to react and this cause to decrease resulting in the
yield of photodegradation decreased at the pH = 11. At pHs < 9, the g-C3N4 NTs surface has a positive
charge and the tetracycline molecules also have a positive charge due to protonation. Therefore, the
electrostatic force between the photocatalyst and the tetracycline is the repulsive force, which leads to a
decrease in degradation efficiency. Although, the highest percentage of pesticide degradation was
obtained at an alkaline pH but, due to the economic aspect and ease of operation treatment at the neutral

Page 10/21
pH, investigation of other parameters in photocatalytic degradation of the mentioned pesticides was
performed at pH = 7

3.3.5. Effect of chemical scavengers


To probe role of reactive species in the photodegradation of tetracycline in the present Ru(II) complex/g-
C3N4 NTs, trapping tests were investigated under optimal condonation. For this purpose, isopropyl
alcohol as a hydroxyl radical scavenger, ethylenediamine tetra acetic acid as a hole scavenger and
sodium sulphate as an electron scavenger were added separately to the reaction mixture and result
compared to no scavenger addition. In all three cases, the addition of scavengers reduced the rate and
degradation percentage. According to these results, it can be deduced that the hydroxyl radical (•OH) and
species generated via hole carriers as key drivers are involved in the photodegradation of tetracycline.

3.3.6. Investigation of synergistic effect


In order to study the synergistic effect, tetracycline degradation reaction in the presence of g-C3N4 NTs
and Ru(II) complex/g-C3N4 NTs was investigated separately and the results are reported in Table1.
According to the results, g-C3N4 nanotubes alone are able to destroy 62% of tetracycline in 90 minutes.
After loading the ruthenium complex, the degradation percentage was higher than 99% for the same
period. It seems combining two photon-active substances; g-C3N4 NTs and Ru(II) complex has been able
to optimize the prepared photocatalyst bandgap. Therefore, the synergistic effect of Ru(II) complex and g-
C3N4 NTs enhances the photocatalytic properties. 

 
Table 1
The kinetic parameters of photodegradation of tetracycline in the
present of different catalyst components
Entry   Tetracycline

Photocatalyst components Degradation(%) k (1/min)

1 g-C3N4 NTs 62 0.0109

2 Ru (II) complex/g-C3N4 NTs 99 0.0417

Reaction condithion: tetracycline concentration: 10 mg/L; photocatalyst dosage: 30 mg; irradiation time:
90 min; light intensity: 60 Watt; temperature: 25°C; pH:7

4. Conclusions
In this study, Ru (II) comples/g- C3N4 nanocatalyst was synthesized and its photocatalytic properties for
the degradation of tetracycline antibiotic was investigated. The synthesized photocatalyst were evaluated
by various methods such as Infrared spectroscopy, X-ray diffraction, scanning and transmission electron
microscopy, and visible ultraviolet spectroscopy. The results showed the successful synthesis of graphite
Page 11/21
nitride carbon nanotubes reinforced with ruthenium (II) complex. In general, one of the most important
properties of a photocatalyst or a catalystis its specific surface area. Therefore, the synthesis of g-C3N4 in
the form of nanotubes had a significant performance in photocatalytic properties. In addition to
increasing the specific surface area, the accessibility surface for light also increases automatically.
Finally, since the bulk graphite carbon nitride alone is a suitable option for photocatalytic degradation of
pollutants, the use of improved its, which has a higher specific surface area and better activity, is not far-
fetched.

Declarations
Acknowledgments Thanks are due to the Iranian Nanotechnology Initiative and the research council of
Shahid Bahonar university of Kerman and chemical engineering department for supporting of this work. 

Data availability statement

The data that support the findings of this study are available on request from the corresponding author.
The data are not publicly available due to privacy or ethical restrictions

References
1. I. Chopra, M. Roberts. Tetracycline antibiotics: mode of action, applications, molecular biology, and
epidemiology of bacterial resistance. Microbiology and molecular biology reviews. 2001 Jun
1;65(2):232 – 60
2. M. Conde-Cid, A. Núñez-Delgado, M.J. Fernández-Sanjurjo, E. Álvarez-Rodríguez, D. Fernández-
Calviño, Arias-Estévez M. Tetracycline and Sulfonamide Antibiotics in Soils: Presence, Fate and
Environmental Risks. Processes. 8(11), 1479 (2020 Nov)
3. A. Fiaz, D. Zhu, J. Sun, Environmental fate of tetracycline antibiotics: degradation pathway
mechanisms, challenges, and perspectives. Environ. Sci. Europe 33(1), 1–7 (2021 Dec)
4. R.A. Smith, N.M. M’ikanatha, A.F. Read, Antibiotic resistance: a primer and call to action. Health
communication. 2015 Mar 4;30(3):309–14
5. B. Zhu, Z. Zong, X. Zhang, D. Zhang, L. Cui, C. Bi, Y. Fan, Highly Selective and Stable Zn (II)-Based
Metal–Organic Frameworks for the Detections of Tetracycline Antibiotic and Acetone in Aqueous
System. Appl. Organomet. Chem. 34(7), e5518 (2020 Jul)
6. S. Hariganesh, S. Vadivel, D. Maruthamani, M. Kumaravel, B. Paul, N. Balasubramanian, T.
Vijayaraghavan, Facile large scale synthesis of CuCr2O4/CuO nanocomposite using MOF route for
photocatalytic degradation of methylene blue and tetracycline under visible light. Appl. Organomet.
Chem. 34(2), e5365 (2020 Feb)
7. H. Park, Y.K. Choung, Degradation of antibiotics (tetracycline, sulfathiazole, ampicillin) using
enzymes of glutathion S-transferase. Human and Ecological Risk Assessment: An International
Journal. 2007 Sep 18;13(5):1147–55

Page 12/21
8. V. Homem, L. Santos. Degradation and removal methods of antibiotics from aqueous matrices–a
review. Journal of environmental management. 2011 Oct 1;92(10):2304-47
9. O.M. Rodriguez-Narvaez, J.M. Peralta-Hernandez, A. Goonetilleke, E.R. Bandala, Treatment
technologies for emerging contaminants in water: A review. Chem. Eng. J. 1, 323:361–380 (2017
Sep)
10. J. Xue, C. Huang, Y. Zong, J. Gu, M. Wang, S. Ma, Fe (III)-grafted Bi2MoO6 nanoplates for enhanced
photocatalytic activities on tetracycline degradation and HMF oxidation. Appl. Organomet. Chem.
33(11), e5187 (2019 Nov)
11. F. Al Marzouqi, N.A. Al-Balushi, A.T. Kuvarega, S. Karthikeyan, R. Selvaraj, Thermal and hydrothermal
synthesis of WO3 nanostructure and its optical and photocatalytic properties for the degradation of
Cephalexin and Nizatidine in aqueous solution. Mater. Sci. Engineering: B 1, 264:114991 (2021 Feb)
12. M.G. Bhosale, R.S. Sutar, S.S. Londhe, M.K. Patil, Sol–gel method synthesized Ce-doped TiO2 visible
light photocatalyst for degradation of organic pollutants. Applied Organometallic Chemistry.:e6586
13. Y.X. Zhang, Y. Jia, Synthesis of MgO/TiO2/Ag composites with good adsorption combined with
photodegradation properties. Mater. Sci. Engineering: B 1, 228:123–131 (2018 Feb)
14. H.V. Vasei, S.M. Masoudpanah, M. Adeli, M.R. Aboutalebi, M. Habibollahzadeh, Mesoporous
honeycomb-like ZnO as ultraviolet photocatalyst synthesized via solution combustion method.
Mater. Res. Bull. 1, 117:72–77 (2019 Sep)
15. M. Zhu, H. Chen, Y. Dai, X. Wu, Z. Han, Y. Zhu, Novel n-p‐n heterojunction of AgI/BiOI/UiO‐66
composites with boosting visible light photocatalytic activities. Appl. Organomet. Chem. 35(5),
e6186 (2021 May)
16. C. Zhang, Y. Zhou, Y. Zhang, S. Zhao, J. Fang, X. Sheng, The investigation of Ag decorated double-
wall hollow TiO2 spheres as photocatalyst. Appl. Organomet. Chem. 32(3), e4160 (2018 Mar)
17. M. Zangiabadi, T. Shamspur, A. Saljooqi, A. Mostafavi, Evaluating the efficiency of the GO-
Fe3O4/TiO2 mesoporous photocatalyst for degradation of chlorpyrifos pesticide under visible light
irradiation. Appl. Organomet. Chem. 33(5), e4813 (2019 May)
18. N.P. de Moraes, C.M. Goes, D.C. Sperandio, R. da Silva Rocha, R. Landers, T. Paramasivam, L.A.
Rodrigues, Development of a new zinc oxide/tin oxide/carbon xerogel photocatalyst for visible light
photodegradation of 4-chlorophenol. Mater. Sci. Engineering: B 1, 269:115183 (2021 Jul)
19. G. Dong, Y. Zhang, Q. Pan, J. Qiu, A fantastic graphitic carbon nitride (g-C3N4) material: electronic
structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol., C 1, 20:33–50
(2014 Sep)
20. R. Zhang, X. Zhang, S. Liu, J. Tong, F. Kong, N. Sun, X. Han, Y. Zhang, Enhanced photocatalytic
activity and optical response mechanism of porous graphitic carbon nitride (g-C3N4) nanosheets.
Mater. Res. Bull. 1, 140:111263 (2021 Aug)
21. Z. Cui, H. Yang, X. Zhao, Enhanced photocatalytic performance of g-C3N4/Bi4Ti3O12 heterojunction
nanocomposites. Mater. Sci. Engineering: B 1, 229:160–172 (2018 Mar)

Page 13/21
22. H. Shi, J. Fu, W. Jiang, Y. Wang, B. Liu, J. Liu, H. Ji, W. Wang, Z. Chen, Construction of g-
C3N4/Bi4Ti3O12 hollow nanofibers with highly efficient visible-light-driven photocatalytic
performance. Colloids Surf., A 20, 615:126063 (2021 Apr)
23. G. Gebreslassie, P. Bharali, U. Chandra, A. Sergawie, P.K. Baruah, M.R. Das, E. Alemayehu,
Hydrothermal synthesis of g-C3N4/NiFe2O4 nanocomposite and its enhanced photocatalytic
activity. Appl. Organomet. Chem. 33(8), e5002 (2019 Aug)
24. Z. Chen, S. Zhang, Y. Liu, N.S. Alharbi, S.O. Rabah, S. Wang, X. Wang, Synthesis and fabrication of g-
C3N4-based materials and their application in elimination of pollutants. Sci. Total Environ. 5, 139054
(2020 May)
25. M. Singh, M. Goyal, K. Devlal. Size and shape effects on the band gap of semiconductor compound
nanomaterials. Journal of Taibah University for Science. 2018 Jul 4;12(4):470- 5
26. K.C. Devarayapalli, S.P. Vattikuti, T.V. Sreekanth, K.S. Yoo, P.C. Nagajyothi, J. Shim, Hydrogen
production and photocatalytic activity of g-C3N4/Co‐MOF (ZIF‐67) nanocomposite under visible light
irradiation. Appl. Organomet. Chem. 34(3), e5376 (2020 Mar)
27. C.P. Goh, P.E. Lim, Potassium permanganate as oxidant in the COD test for saline water samples.
ASEAN J. Sci. Technol. Dev. 25(2), 383–393 (2008)
28. Q. Wang, P. Li, Z. Zhang, C. Jiang, K. Zuojiao, J. Liu, Y. Wang. Kinetics and mechanism insights into
the photodegradation of tetracycline hydrochloride and ofloxacin mixed antibiotics with the flower-
like BiOCl/TiO2 heterojunction. Journal of Photochemistry and Photobiology A: Chemistry. 2019 Jun
1;378:114 – 24
29. J. He, Y. Zhang, Y. Guo, G. Rhodes, J. Yeom, H. Li, W. Zhang. Photocatalytic degradation of
cephalexin by ZnO nanowires under simulated sunlight: Kinetics, influencing factors, and
mechanisms. Environment international. 2019 Nov 1;132:105105
30. F. Wang, Y. Feng, P. Chen, Y. Wang, Y. Su, Q. Zhang, Y. Zeng, Z. Xie, H. Liu, Y. Liu, W. Lv, Photocatalytic
degradation of fluoroquinolone antibiotics using ordered mesoporous gC3N4 under simulated
sunlight irradiation: kinetics, mechanism, and antibacterial activity elimination. Applied Catalysis B:
Environmental. 2018 Jul 5;227:114–22
31. A. Thomas, A. Fischer, F. Goettmann, M. Antonietti, J.O. Müller, R. Schlögl, J.M. Carlsson, Graphitic
carbon nitride materials: variation of structure and morphology and their use as metal-free catalysts.
J. Mater. Chem. 18(41), 4893–4908 (2008)
32. E. Alwin, K. Kočí, R. Wojcieszak, M. Zieliński, M. Edelmannová, M. Pietrowski, Influence of high
temperature synthesis on the structure of graphitic carbon nitride and its hydrogen generation ability.
Materials. 13(12), 2756 (2020 Jan)
33. H. Zhao, X.L. Chen, C. Jia, T. Zhou, X. Qu, J. Jian, Y. Xu, T. Zhou, WITHDRAWN: A facile
mechanochemical way to prepare g-C3N4. Materials Science and Engineering: B. 2005 Sep
25;122(3):226–30
34. Z. Jin, Q. Zhang, S. Yuan, T. Ohno, Synthesis high specific surface area nanotube gC 3 N 4 with two-
step condensation treatment of melamine to enhance photocatalysis properties. RSC Adv. 5(6),
Page 14/21
4026–4029 (2015)
35. C. Hu, Y.C. Chu, M.S. Wang, X.H. Wu, Rapid synthesis of g-C3N4 spheres using microwave-assisted
solvothermal method for enhanced photocatalytic activity. Journal of Photochemistry and
Photobiology A: Chemistry. 2017 Nov 1;348:8–17
36. Z. Chen, S. Zhang, Y. Liu, N.S. Alharbi, S.O. Rabah, S. Wang, X. Wang, Synthesis and fabrication of g-
C3N4-based materials and their application in elimination of pollutants. Sci. Total Environ. 5, 139054
(2020 May)
37. M. Ismael, Y. Wu, A mini-review on the synthesis and structural modification of gC3N4 based
materials, and their applications in solar energy conversion and environmental remediation.
Sustainable Energy & Fuels 3(11), 2907–2925 (2019)
38. S.C. Yan, Z.S. Li, Z.G. Zou. Photodegradation performance of g-C3N4 fabricated by directly heating
melamine. Langmuir. 2009 Sep 1;25(17):10397-401
39. X. Zhao, Y. Zhang, X. Zhao, X. Wang, Y. Zhao, H. Tan, H. Zhu, W. Ho, H. Sun, Y. Li, Urea and melamine
formaldehyde resin-derived tubular g-C3N4 with highly efficient photocatalytic performance. ACS
applied materials & interfaces. 2019 Jul 18;11(31):27934–43
40. G. Goglio, D. Foy, G. Demazeau, State of Art and recent trends in bulk carbon nitrides synthesis.
Materials Science and Engineering: R: Reports. 2008 Jan 7;58(6):195–227
41. Q. Gan, W. Shi, Y. Xing, Y. Hou, A polyoxoniobate/g-C3N4 nanoporous material with high adsorption
capacity of methylene blue from aqueous solution. Frontiers in chemistry. 2018 Jan 31;6:7
42. X. Bai, L. Wang, R. Zong, Y. Zhu, Photocatalytic activity enhanced via g-C3N4 nanoplates to
nanorods. J. Phys. Chem. C 16(19), 9952–9961 (2013 May) 117(
43. M. Yin, F. Jia, C. Wu, P. Zheng, Y. Fan, Z. Li, Coupling g-C3N4 nanobelts and Cu (OH) 2 nanoparticles
with TiO2 for visible-light photocatalytic H2 production. Mater. Sci. Engineering: B 1, 223:35–42
(2017 Sep)
44. A. Fischer. " Reactive hard templating": from carbon nitrides to metal nitrides (Doctoral dissertation,
Universität Potsdam)
45. X. Wang, L. Sø, R. Su, S. Wendt, P. Hald, A. Mamakhel, C. Yang, Y. Huang, B.B. Iversen, F. Besenbacher.
The influence of crystallite size and crystallinity of anatase nanoparticles on the photo-degradation
of phenol. Journal of catalysis. 2014 Feb 1;310:100-8
46. Y. Chen, B. Yang, W. Xie, X. Zhao, Z. Wang, X. Su, C. Yang. Combined soft templating with thermal
exfoliation toward synthesis of porous g-C3N4 nanosheets for improved photocatalytic hydrogen
evolution. journal of materials research and technology. 2021 Jul 1;13:301 – 10
47. R. Rastogi, R. Kaushal, S.K. Tripathi, A.L. Sharma, I. Kaur, L.M. Bharadwaj, Comparative study of
carbon nanotube dispersion using surfactants. Journal of colloid and interface science. 2008 Dec
15;328(2):421–8
48. V. Sangeetha, N. Kanagathara, R. Sumathi, N. Sivakumar, G. Anbalagan, Spectral and thermal
degradation of melamine cyanurate. Journal of Materials. 2013;262094

Page 15/21
49. D. Luo, W. Duan, Y. Liu, N. Chen, Q. Wang, Melamine cyanurate surface treated by nylon of low
molecular weight to prepare flame- retardant polyamide 66 with high flowability. Fire Mater. 43(3),
323–331 (2019 Apr)
50. G. Xin, Y. Meng, Pyrolysis synthesized g-C3N4 for photocatalytic degradation of methylene blue.
Journal of Chemistry. 2013 Jan 1;2013
51. X. Ren, X. Meng, J. Ren, F. Tang, Graphitic carbon nitride nanosheets with tunable optical properties
and their superoxide dismutase mimetic ability. RSC Adv. 6(95), 92839–92844 (2016)
52. A. Kumar, P. Kumar, C. Joshi, M. Manchanda, R. Boukherroub, S.L. Jain, Nickel decorated on
phosphorous-doped carbon nitride as an efficient photocatalyst for reduction of nitrobenzenes.
Nanomaterials. 6(4), 59 (2016 Apr)
53. C. Liu, X. Dong, Y. Hao, X. Wang, H. Ma, X. Zhang, A novel supramolecular preorganization route for
improving gC 3 N 4/gC 3 N 4 metal-free homojunction photocatalysis. New J. Chem. 41(20), 11872–
11880 (2017)
54. S.P. Pattnaik, A. Behera, S. Martha, R. Acharya, K. Parida, Facile synthesis of exfoliated graphitic
carbon nitride for photocatalytic degradation of ciprofloxacin under solar irradiation. J. Mater. Sci.
54(7), 5726–5742 (2019 Apr)
55. F. Jensen, Activation energies and the Arrhenius equation. Qual. Reliab. Eng. Int. 1(1), 13–17 (1985
Jan)
56. B. Zhu, P. Xia, W. Ho, J. Yu, Isoelectric point and adsorption activity of porous g-C3N4. Appl. Surf. Sci.
30, 344:188–195 (2015 Jul)

Figures

Figure 1

FT-IR spectra of melamine-cyanuric acid (a), g-C3N4 NTs (b), and Ru(II)complex/g-C3N4 NTs (c).

Page 16/21
Figure 2

The spectrum of a) UV-Vis spectrum b) curve of (αhν)2 versus hν of g-C3N4 NTs to bandgap estimation.

Page 17/21
Figure 3

XRD patterns of a) Melamine-Cyanuric acid , b) Ru(II)complex/g-C3N4 NTs

Figure 4

Scanning electron microscopy (SEM) photographs (a), energy‐dispersive X‐ray pattern (b) and element
distribution map (c) of Ru(II)complex/g-C3N4 NTs photocatalyst.

Page 18/21
Figure 5

TEM images of Ru(II)complex/g-C3N4 NTs photocatalyst.

Figure 6

a) Influence of temperature on the photocatalytic degradation of tetracycline, b) Plot of ln(C/C0) versus


time to investigation of temperature effect on the degradation rate. Reaction conditions: drug
concentration: 10 mg/L, irradiation time: 90 min, light intensity: 60 Watt, temperature: 25°C, pH: 7

Figure 7

Influence of irradiation time on the photocatalytic degradation of tetracycline in water. Drug


concentration: 10 mg/L, Ru(II) complex/g-C3N4 NTs amount: 30 mg, light intensity: 60 Watt, temperature:
25 °C, pH: 7

Page 19/21
Figure 8

a) Influence of temperature on the photocatalytic degradation of tetracycline in the first 30 minutes, b)


Plot of ln(C/C0) versus time to investigation of temperature effect on the degradation rate. Reaction
conditions: drug concentration: 10 mg/L, Ru(II) complex/g-C3N4 NTs amount: 30 mg, irradiation time: 90
min, light intensity: 60 Watt, pH: 7

Figure 9

Page 20/21
Plot of Ln k versus 1/T to estimation of Ea.

Figure 10

a) Influence of pH on the photocatalytic degradation of tetracycline in the first 30 minutes, b) Plot of


ln(C/C0) versus time to investigation of pH effect on the degradation rate. Reaction conditions: drug
concentration: 10 mg/L, Ru(II) complex/g-C3N4 NTs amount: 30 mg, irradiation time: 90 min, light
intensity: 60 Watt, Temperature: 25°C

Page 21/21

You might also like