You are on page 1of 11

Journal of Environmental Chemical Engineering 7 (2019) 103397

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Photocatalytic treatment of a multicomponent petrochemical wastewater by T


floatable ZnO/Oak charcoal composite: Optimization of operating
parameters
N. Tafreshia, S. Sharifniab, , S. Moradi Dehaghia

a
Department of Chemistry, Tehran North Branch, Islamic Azad University, Tehran, Iran
b
Catal. Res. Cen., Dept. Chem. Eng., Razi Univ., Kermanshah 67149-67246, Iran

ARTICLE INFO ABSTRACT

Keywords: This paper aims to investigate the potential of photocatalytic processes for the degradation of multicomponent
Photocatalyst industrial wastewaters. Commercial ZnO nanoparticles were immobilized on pieces of carbonized Oak as a
Industrial ammonia wastewater floatable support and it was applied for the decontamination of industrial ammonia wastewater from a petro-
Floating support chemical plant under UV irradiation. Wastewater characterization confirmed the presence of ammonia, BOD,
Box-Behnken
COD, CaCO3, Cl− and SO42−. Box-Behnken experimental design was used to find out the best experimental
conditions and the maximum ammonia removal. A second-order polynomial model in terms of pH, catalyst
dosage and light intensity was suggested for prediction of the maximum efficiency of ammonia removal.
Accuracy of the model was verified by p-value of the model (< 0.0001), p-value of lack of fit (0.3987), R2
(0.9854) and adjusted R2 (0.9666). This model suggested that light intensity is the most effective parameter
maximizing the ammonia removal. The optimum conditions under which the maximum ammonia removal of
86.7% was observed are as follows: pH 9, catalyst dosage of 2 g/L, and light intensity of 250 W. This efficiency is
close to the predicted value of 93%. The other contaminants present in the wastewater were also decomposed
significantly by ZnO/Oak charcoal composite. This study can be useful for who are working on the real appli-
cations of photocatalysis processes.

1. Introduction ammonia wastewater poses a great threat to the aquatic organisms and
must be treated before discharging to the environment.
Ammonia wastewater is one of the main destructive sewage and has Wastewater treatment refers to any type of physical, chemical and
the detrimental impacts on the communities and ecosystems. This ni- biological methods that reduce the concentration of hazardous com-
trogen contaminant must be removed from wastewaters prior to the pounds in aqueous media either by separation of them (like adsorption
reuse or release into the environment. Most of quality standards of the [5] and membrane based [6] methods) or by the breakdown of complex
treated wastewater specified the maximum ammonia concentration of organics into simpler compounds that are stable and safer such as ad-
0.5 mg/L for the environment safety purpose [1]. Depending on the pH vanced oxidation processes (AOPs) [7–9]. The most widely being used
and temperature of the water, two forms of inorganic nitrogen species, remediation processes of ammonia wastewater in high concentration
NH3 (free ammonia) and NH4+ (ionized ammonia), are in equilibrium are air or steam stripping [10–12], ionic exchange [13,14], breakpoint
in an ammonia wastewater. The free ammonia species is gaseous and chlorination [15,16], electrodialysis [17,18] and nitrification-deni-
can be stripped completely above pH 12, while its ionized form is so- trification [19,20] processes. However, these processes cannot com-
luble in water and remains as soluble ammonia ions entirely below pH 7 pletely remove ammonia from aqueous media. Among these processes,
[2,3]. In the real ammonia wastewaters, nitrogen is also oxidized bio- ammonia stripping technique is a highly efficient way of treating am-
logically by nitrifying bacteria, so the biochemical oxygen demand monia wastewater. It was reported that ammonia stripping method
(BOD) is expected to have a high value. Other expectable constituents could remove ammonia with the efficiency as much as 90% [21]. Al-
of real ammonia wastewaters are chemical oxygen demand (COD), total though ammonia stripping is a low-cost and effective procedure, it has
suspense solid (TSS), CaCO3, SO42−, etc. [4]. Thus, the raw industrial some disadvantages like fouling problems [22], sludge production [23]


Corresponding author.
E-mail address: sharif@razi.ac.ir (S. Sharifnia).

https://doi.org/10.1016/j.jece.2019.103397
Received 24 May 2019; Received in revised form 7 August 2019; Accepted 30 August 2019
Available online 11 September 2019
2213-3437/ © 2019 Elsevier Ltd. All rights reserved.
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Scheme 1. Schematic diagram of photocatalyst preparation, the rectangles show the catalyst preparation steps and the ellipses show the final products.

Table 1 processes for the decontamination of effluents generated from various


Factors and levels in Box-Behnken design. types of industry sectors like petrochemical plants [25–27]. One ad-
Factor Coded Levles vantage of photocatalytic degradation of ammonia wastewater is the
capability of this process to be scaled up from the laboratory scale to
Low (-1.000) Medium (0.000) High (1.000) practical applications [28]. Apart from photogenerated holes, free hy-
droxyl and superoxide radicals are powerful oxidizing agents capable of
pH (X1) 4 8 12
Catalyst dosage (g/L) (X2) 0.5 1.25 2
degradation of ammonia and other organic and inorganic contaminants
Light intensity (w) (X3) 80 125 250 [29]. ZnO is a wide band gap photocatalyst with n-type conductivity
and high electron mobility of 205 cm2/V s at room temperature [30].
ZnO have shown good stability and photocatalytic activity in various
and needing to NH3 absorption unit as a secondary process [24] to aqueous media [31,32]. The redox potential of its valence band holes is
prevent additional environmental issues caused by NH3 releasing to the + 2.57 V (vs. SHE) which completely meets the potential required for
environment. All the other methods have their own favorable char- oxidizing the pollutants while also producing %OH species (Eº =
acteristics, but each has its limitations like formation of secondary +2.81 V vs. SHE). Similarly, the redox potential of its conduction band
pollutants, needing to high cost equipment and limitation of applic- electrons is -0.53 V (vs. SHE) which is completely in accordance to the
ability, especially when a real wastewater is targeted [10]. For these potential needed to form %O2− (Eº = +0.89 V vs SHE) [27].
reasons, AOPs have been emerged as alternative to or auxiliary process One of the main limitations of the photocatalytic processes in liquid
of the traditional methods since they are able to deal with the problem systems is the ineffective penetration of light through the wastewater
of ammonia degradation with high efficiency. for activating the settled photocatalyst particles by photon absorption.
Photocatalytic wastewater treatment is one of the oxidation For the real wastewaters which have inherently turbidity, the

2
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Table 2 impregnation and exposed to UV light irradiation to decompose the


The plan of experiments along with the experimental and predicted values for pollutants of industrial ammonia wastewater discharged from Ker-
Ammomia photodegradation %. manshah Petrochemical Company, Iran. The main parameters of pH,
Run Coded Levles Ammonia photodegradation (%) catalyst dosage, and light intensity were optimized by Box-Behnken
approach, which is an affordable and time-saving design [42,43], to
X1 X2 X3 Experimental Predicted maximize the ammonia removal efficiency. This study can be con-
sidered for the researchers who focus on the actual usages of photo-
1 0.000 0.000 0.000 84.00 84.64
2 1.000 −1.000 0.000 78.00 78.44 catalytic processes since there have been only a few reports dealing
3 1.000 0.000 −1.000 71.50 70.68 with the photocatalytic treatment of real ammonia effluents.
4 0.000 0.000 0.000 85.50 84.64
5 1.000 0.000 1.000 89.70 90.90
2. Material & method
6 0.000 0.000 0.000 85.90 84.64
7 0.000 0.000 0.000 85.30 84.64
8 −1.000 0.000 −1.000 65.00 67.55 2.1. Material
9 1.000 1.000 0.000 86.00 85.19
10 0.000 1.000 −1.000 76.00 75.01 ZnO powder, ethanol, and nitric acid were prepared from Merck Co.
11 −1.000 1.000 0.000 82.00 82.06
The Oak wood was obtained from Kermanshah’s forests, Iran, as the
12 −1.000 0.000 1.000 90.70 87.78
13 0.000 −1.000 −1.000 69.00 68.26 source of Oak charcoal. The used double distilled water was produced
14 0.000 1.000 1.000 93.50 92.24 in our lab. We applied industrial ammonia wastewater obtained from
15 −1.000 −1.000 0.000 75.00 75.31 Kermanshah Petrochemical Co., Iran.
16 0.000 0.000 0.000 82.50 84.64
17 0.000 −1.000 1.000 88.50 88.48
2.2. Photocatalyst preparation

The chopped Oak wood with average size of


(6 mm × 4 mm × 2 mm) first was soaked in 2 mol/L HNO3 at 60 °C to
remove the likely metal ions for 12 h. To clean up the support pieces
from the pigments inside its tissues, they were plunged into absolute
alcohol and deionized water at 60 °C for 12 h. Finally, they were dried
at 60 °C for 12 h. The dried wood pieces were heated up to 450 °C for
carbonization. A slurry composed of zinc oxide powder, ethanol, and
dilute nitric acid was prepared and sonicated for 30 min. ZnO nano-
particles were impregnated on the support surface by soaking the dried
support pieces into the prepared slurry and agitation until ethanol was
completely evaporated. The products then were dried at 120 °C for 12 h,
followed by calcination at 450 °C for 2 h under an inert atmosphere of
nitrogen. Scheme 1 shows the tutorial of the preparation of photo-
catalyst composites.

2.3. Methods of characterization

Fig. 1. XRD patterns of ZnO and ZnO/Oak charcoal samples. The prepared samples was characterized by XRD (XRD, PHILIPS
PW1140) to study the structure of the samples. The morphology, shape
penetration of light is even more expected. In the laboratory scale, this and size of the samples were found by FESEM analysis (FESEM, MIRA3-
issue can be addressed by agitation which is not recommended for the TESCAN). The optical property of the samples was measured by UV–vis
real applications. The use of floatable supports on which the photo- DRS spectroscopy (Jasco V-670 spectrophotometer, Japan). The texture
catalyst particles are loaded is a suitable way for the practical appli- properties of the photocatalysts were measured by BET method. And
cations of photocatalytic wastewater treatment [4,33]. Some floatable the pore size distribution was investigated by BJH method.
supports of photocatalysts are minerals like LECA [4,34] and perlite The concentration of ammonia before and after photocatalysis was
[27,35], exfoliated vermiculite (EV) [36], hollow glass microbeads measured by UV spectrophotometer. Using an American public health
[37], foams and polymers such as polyurethane foams [38], polystyrene association standards (APHA, 1998), the concentration of other com-
[39,40], etc. ponent present in the industrial wastewater was measured as follows.
Using low-cost, achievable photocatalyst supports with simple The concentration of COD was determined by a Hach-2000 spectro-
synthesis procedure is essential and must be considered for the photo- photometer using a dichromate solution as the oxidant in strong acid
catalytic remediation of large scale wastewaters. In the previous study media via closed reflux, colorimetric method through APHA, 1995
[41] we investigated the positive role of the Oak charcoal as support of section 5220-D. Total suspended solids (TSS) were measured by the
commercial ZnO for the degradation of model ammonia wastewater. gravimetric method via APHA, 1995 section 2540-D. BOD5 was ana-
The photocatalytic efficiency of 43% was obtained by ZnO nano- lyzed by using a respiremetry system by 5-day BOD test based on APHA,
particles without using the support which was considerably lower than 1995 section 5210-B. Sulfate (SO4−2) and chloride anions were mea-
80% efficiency observed by ZnO supported on the Oak charcoal. This sured by ion exchange chromatography (Dionex, Model: DX-120) after
indicates that floating of photocatalysts is a key way of boosting pho- filtering the samples through a 0.45 mm filter.
tocatalytic activity. Therefore, the importance of the study presented
here is: (I) the assessment of floating effect of photocatalyst particles in 2.4. Photoreactor tests
the photocatalytic application using low-cost, available materials and
(II) investigation of potential of photocatalytic process for the decom- The photodegradation process was carried out in a Pyrex glass
position of real ammonia wastewater. To this purpose, commercial ZnO vessel with the inner volume of 1.5 L which was top-irradiated by an Hg
nanoparticles were loaded on the low-density Oak charcoal by lamp (maximum wavelength of 254 nm). The inside of the experimental
set-up was completely shielded by a metal foil to avoid losing light.

3
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Fig. 2. FESEM images of (a) commercial ZnO nanoparticles, (b) and (c) Oak charcoal, and (d) a piece of Oak charcoal coated by ZnO.

ZnO/Oak charcoal composite (10 wt.% ZnO) was floated on 1.0 L of real determined range of the input variables. In the work presented here, by
ammonia wastewater and then irradiated by UV light for 4 h. pH of the applying Box-Behnken experimental design (Design Expert software
solution was adjusted by 0.1 N HCl and/or NaOH solutions. An air (Ver. 7.0.0)), the removal efficiency of ammonia from real wastewater
diffuser on the bottom of the reactor supplied the oxygen needed for the as the output variable by ZnO/Oak charcoal was modeled mathemati-
photocatalytic reactions. The temperature of reactor was kept constant cally in terms of three main factors of pH (X1), catalyst dosage (X2), and
at 20 °C by circulating cooling water. Prior to the photocatalysis tests, light intensity (X3). Three levels of the input variables are given in
the adsorption capacity of the composite photocatalysts was measured Table 1. A set of 17 experiments in three levels (with three times re-
by a blank test, a non-irradiated test for 120 min. After irradiation, the plication of the center point) were designed to compare the differences
content of the photoreactor was sampled to determine the concentra- between the observed and predicted values. Table 2 presents the
tion of ammonia and other contaminant based on the measurement schedule of experiments and the predicted and observed responses. The
methods mentioned in the previous section. experimental data were fitted well by a second-order polynomial
mathematic model which is shown in Eq. (1).
k k k k
2.5. Experimental design and statistical analysis
Y = b0 + bi x i + bii x i2 + bij x i xj +
i=1 i=1 i=1 j> 1 (1)
Experimental design systematically examines the importance of the
independent variables (inputs) and their interactions and finally max- where Y is the modeled output, b0 is the offset term. bi, bii, and bij are
imizes or minimizes the dependent variables (outputs) in the the coefficient of linear, squared, and interaction effects, respectively. ε

4
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Fig. 3. N2 adsorption-desorption isotherms and pore size distributions of Oak charcoal and ZnO/charcoal samples.

Table 3 Table 4
The textural properties of the Oak charcoal and ZnO/charcoal. Results of ANOVA.
Sample Surface areaa (m2/g) Pore volumeb (cm3/g) Pore diameterc (nm) Source Sum of Degree of Mean F-value p-value
squares freedom square prob > F
Oak charcoal 5.233 0.074637 21.21
ZnO/charcoal 15.063 0.002775 1.98 Model 1020.17 9 113.35 52.51 < 0.0001
X1 19.53 1 19.53 9.05 0.0197
a
BET specific surface area. X2 91.13 1 91.13 42.22 0.0003
b
measured at P/Pº = 0.990. X3 818.10 1 818.10 379.01 < 0.0001
c
calculated from the desorption branch of the isotherm using the BJH X1 X2 0.25 1 0.25 0.12 0.7436
X1 X3 14.06 1 14.06 6.51 0.0380
method.
X2 X3 1.00 1 1.00 0.46 0.5180
X12 50.33 1 50.33 23.32 0.0019
X22 3.66 1 3.66 1.70 0.2340
X32 16.13 1 16.13 7.47 0.0292
Lack of Fit 7.36 3 2.45 1.27 0.3987
Pure Error 7.75 4 1.94
Residual 15.11 7 2.16
Total 1035.28 16

R2 = 0.9854, Adj-R2 = 0.9666, Adeq precision = 25.792, Std. Dev. = 1.47, C.V.
% = 1.8.

Table 5
Characterization of various contaminants in the industrial ammonia wastewater
before and after photocatalytic treatment.
Component Unite Initial value Final value Conversion (%)

pH − 9.8 10.2 −
Total hardness ppm 440 205 53.4
CaCO3 ppm 75 21 72.0
Ammonia ppm 310 41 86.7
Chloride ppm 1280 915 28.5
Fig. 4. Room temperature UV–vis spectra of ZnO and ZnO/Oak charcoal sam- Sulfate ppm 320 168 46.0
ples. BOD ppm 75 48 36.0
COD ppm 1610 674 58.0

is the random error.Also, i and j are the index number, and xi stands for
the dimensionless codded variable of Xi. several groups are equal. By ANOVA, the impact of linear, interactive,
The analysis of variance (ANOVA), used in the paper presented, is a and quadratic effects of variables can be determine. Regression coeffi-
powerful statistical method revealing whether or not the means of cients of adjusted and predicted (R2-adj and R2), lack of fit (LOF), and

5
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Fig. 5. (a) Residual plot of the predicted and actual values, (b) normal probability of internally studentized residuals.

degrees of freedom are other statistical tools measuring the significance this property also make it an appealing support for the actual waste-
of the model applied. water treatment because of the possibility of adsorbing pollutant mo-
lecules inside the pores and cavities. Fig. 2 (d) shows a piece of Oak
charcoal coated with ZnO particles. As can be seen, the larger pores
3. Results & discussion
were not completely blocked by the photocatalyst.
Fig. 3 shows N2 adsorption-desorption analysis and pore size dis-
3.1. Results of photocatalyst characterization
tribution of the Oak charcoal and ZnO/charcoal samples. The isotherm
of the Oak charcoal may be classified as type II, meanwhile the ZnO/
Fig. 1 shows the XRD patterns of ZnO and ZnO/charcoal samples.
charcoal exhibited a type I isotherm. Both samples showed H3 type
The main characteristic peaks at scattering angles of 32, 34, 36, 47, 56,
hysteresis loops. The H3 hysteresis loop is usually appears on solids
62 and 68° can clearly be seen from the pattern of ZnO, relating to the
having aggregated or agglomerated particles with slit-shaped, nonuni-
(100), (002), (101), (102), (110), (103) and (112) facets in the wurtzite
form size pores [44]. Type II isotherm with H3 hysteresis loop is the
structure of ZnO with hexagonal phase, respectively (JCPDS card no.0-
characteristic of nonporous solids or mesoporous and macroporous
3-0888). All the diffraction peaks of ZnO are detected from XRD pattern
powders [45]. This type of isotherm also indicates unrestricted mono-
of the ZnO/charcoal, confirming the presence of ZnO nanoparticles on
layer-multilayer adsorption, multilayer adsorption takes place at high
the support. The broad peak centered at 24° is attributed to the support
P/Pº [46]. The broad hysteresis loop of the Oak charcoal is located at a
which indicates that the Oak charcoal had an amorphous nature.
high relative pressure region (0.79-0.96), which means that there were
To understand the morphology and particle size of the samples,
large pores on it. This result is definitely in agreement with the results
FESEM analysis was applied and the results are given in Fig. 2. There
obtained from FESEM analysis. For the ZnO/charcoal composite, a
are ZnO particles with rod-like shape and particle size distribution of
small hysteresis loop is observed at lower relative pressure (0.34-0.97)
50–200 nm (Fig. 2(a)). According to Fig. 2(b) and (c), Oak charcoal is a
compared to the Oak charcoal sample, indicating the mesoporous fea-
porous support, having a lot of surface spherical cavities with diameters
ture of this catalyst [47]. It is follows that the addition of ZnO leads to
of 5–15 μm. Other than affordability and floatability of Oak charcoal,

6
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Fig. 6. Perturbation plots of pH (A), catalyst dosage (B), and light intensity (C) on the photocatalytic performance of ammonia degradation.

the decrease of pore volume due to the incomplete blockage of cavities significant for p-value lower than 0.05. Accordingly, a prediction model
on the support structure. This assumption was investigated by BJH is suitable provided that LOF is greater than 0.05. Herein, LOF p-value
analysis. From BJH plot of the Oak charcoal, the high peak at 14.18 nm of the model proposed is 0.3987 which is large enough, so the model
indicates the size of mesoporous, and the low and broad peak centered can successfully predict the dependent variable. The p-value of the
at 51.36 nm is ascribed to the larger pores. The BJH plot of the ZnO/ model is < 0.0001, indicating that the model is highly significant. The
charcoal shows two weak peak at 9.18 and 29.76 nm, proving that this predicted R2 (0.9854) and adjusted R2 (0.9666) agree with each other
sample had a mesoporous nature with a narrow pore size distribution. and are very close to 1, demonstrating goodness of fitting the predicted
Table 3 gives the BET parameters of the samples. With loading ZnO model curve on the experimental data. Indeed, the proposed model is
over the Oak charcoal, the pore size and pore volume decreased from unable to explain only 2% of the total variations. The accuracy of the
21.21 to 1.98 nm and 0.074637 to 0.002775 cm3/g while specific sur- model is also confirmed by the predicted versus actual value plot de-
face area increased from 5.233 to 15.063 m2/g. picted in Fig. 5(a). Moreover, Fig. 5(b) indicates that the studentized
For the photocatalysis purposes, optical characterization is im- residuals lie between -2 and +2 without any outliner and the devia-
portant, and thus UV–vis DRS was taken from ZnO and ZnO/charcoal tions of them with the diagonal line are small. The mutual interactions
samples in the wavelength range of 300 to 600 nm (Fig. 4). ZnO is an of X1X2 and X2X3, and quadratic interaction of X22 are insignificant and
excellent UV absorbent mainly in 390 nm. The composite samples can be eliminated from the prediction model. On the other hand, F-
showed better light absorption performance because it can be activated value of 52.51 also revealed the significance of the model. LOF F-value
by photons with lower energy (∼ 460 nm) in visible region. Based on of 1.27 implies that the LOF is not significant relative to the pure error.
Eg = 1245/λa, where Eg and λa respectively are the band gap energy Among the significant variables, the highest F-value and lowest p-value
and the activating wavelength of light, the corresponding band gaps of belong to X3, which means that light intensity is the most important
ZnO and ZnO/charcoal are estimated to be 3.19 and 2.70 eV, respec- parameter affecting the photocatalytic efficiency of ammonia de-
tively. gradation. The significant variables have the effectiveness order of light
intensity > catalyst dosage > pH. Other diagnostic tool confirming
the results of ANOVA is perturbation analysis which remarks the effects
3.2. Nonlinear modeling of significant input variables on the predicted response. From Fig. 6, it
is apparent that the light intensity (C) is the most effective parameter,
pH, catalyst dosage, and light intensity were selected as three key and ammonia degradation would increase with the increment of the
parameters affecting the photocatalytic efficiency of industrial am- light intensity. The two other variables have a negligible effect on the
monia degradation. The experimental matrix design based upon Box- response when they change in the vicinity of the centre point. However,
Behnken method, and the real and predicted responses are tabulated in the effect of catalyst dosage (B) is slightly higher than that of pH (A) as
Table 2. On the basis of the experimental design, the photocatalytic confirmed by p and F tests. Adequate precision, or signal to noise ratio,
ammonia degradation as an output is a function of pH (X1), catalyst is a tool for comparison of the range of the predicted values and the
dosage (X2), and light intensity (X3) by a second-order polynomial average variance at the design points. Adequate precision greater than
equation in terms of coded factors (Eq. (2)). 4 indicates adequate model discrimination. In this particular case it is
25.792, well greater than 4. Very low value of the coefficient of var-
Y = 84.25 + 1.56 X1 + 3.38 X2 + 10.11 X3 – 1.88 X1X3 – 3.51 X12 –
iation, 1.8%, indicates a good reliability of the experimental data.
2.01 X32 (2)

where Y, X1, X2, and X3 stand for efficiency of ammonia removal, pH, 3.3. Response surface analysis
catalyst dosage, and light intensity, respectively. Tables 4 and 5 in-
dicates the results of ANOVA. The independent variables, their inter- The 3D and contour plots showing the effect of mutual interactions
actions, and the regression coefficients can be statistically considered on the response are given in Fig. 7. Fig. 7 (a) and (b) signify that the

7
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

Fig. 7. 3D and cotour plots showing interaction of (a)-(b) pH and catalyst dosage, (c)-(d) pH and light intensity, and (e)-(f) catalyst dosage and light intensity.

photocatalytic efficiency of ammonia degradation can be maximized by maximizing the photocatalytic degradation of ammonia. Promoting
controlling pH and catalyst dosage, the former must be in the basic light intensity accelerates the surface photocatalytic reactions because
range (8–10) and the latter at 2 g/L. It is reasonable that the photo- the number of activating photons per unite area of the photocatalyst
catalytic degradation of ammonia is favorable in basic pH. The presence would increase. In the case of this study, the light intensity of 250 W is
of OH− species from a basic source like NaOH helps more hydroxyl recommended to maximize the response. Fig. 7 (e) and (f), including
radicals to be form by electron capturing from photocatalyst’ CB. Also, the effect of light intensity and catalyst dosage, also verifies the results
in basic medium, the adsorption of ammonia molecules on photo- interpreted above.
catalyst surface increases, for the negatively charged photocatalyst
surface possesses a great affinity with the NH4+ species. Increasing
catalyst dosage corresponds to the increase of both adsorption site and 3.4. Model validation
charge carriers participating in the photocatalytic reactions thus in-
creasing the efficiency of ammonia removal. From Fig. 7 (c) and (d), the Based on the results obtained from the modeling of photocatalytic
effect of concurrent changes in pH and light intensity on the response is performance of ZnO/Oak charcoal, we carried out an experiment in
seen. Clearly, light intensity is the most influential parameter triplicate in the optimum condition, pH 9, catalyst dosage of 2 g/L, and
light intensity of 250 W, to verify the reliability of the model. The

8
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

This study
Ref.

[4]
Removal efficiency

86.7%
96%
Final ammonia concentration

Fig. 8. Results of three times recycling ZnO/charcoal for the photocatalytic


treatment of petrochemical ammonia wastewater.
41
39
250 W high-pressure Hg lamp with maximum wavelength of 254 nm
A comparison between the results of photocatalytic treatment of industrial ammonia wastewater obtained from the present study and our previous work.

Solar light with the mean UV/visible wavelength of 280-390 nm


Light source
Process time (h)

72
4
Catalyst dosage (g/L)

Fig. 9. (a) XRD patterns and (b) FTIR spectra of ZnO/Charcoal before and after
25
2

three times recycling.


pH

11
9

ammonia degradation efficiency of 86.7% was obtained, which is close


Initial ammonia concentration (ppm)

to the predicted value (93%). Other than ammonia, the other con-
taminants were also degraded considerably. The results are tabulated in
Tables 4A and 4B. As can be seen, the photocatalytic treatment sig-
nificantly reduced the concentration of dissolved CaCO3 (72.0%) which
is a main reason of temporary hardness. A comparison between the best
result of ammonia removal obtained here with the result from our
previous work is shown in Table 6.
310
975

3.5. Evaluation of ammonia removal via non-photocatalytic processes


ZnO/charcoal
Photocatalyst

A series of experiments were done at the optimum conditions to


TiO2/LECA

reveal the floating effect of the photocatalyst as well as the participa-


Table 6

tion of other processes in the removal of ammonia from the wastewater.


The quantities of ammonia removal by adsorption (the Oak charcoal in

9
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

darkness without both ZnO and air diffuser), photolysis (UV without industrial ammonia wastewater obtained from a petrochemical plant.
both the photocatalyst and air diffuser), and stripping (via air diffuser ZnO immobilized on Oak charcoal was used as a floatable photocatalyst
(2.0 L/min) without the photocatalyst in darkness) were obtained to be under UV irradiation. Box-Behnken optimization indicated that among
18, 4, and 7%, respectively. These results show that photocatalysis is three parameters of pH, catalyst dosage, and light intensity, light in-
the prevailing process of ammonia removal from real wastewater. tensity is the most effective parameter on the photodegradation of
Besides, ZnO nanoparticles showed the ammonia photodegradation of ammonia. The prediction model suggested the optimum conditions of
37% under UV irradiation, demonstrating the considerable effect of pH 9, catalyst dosage of 2 g/L, and light intensity of 250 W, which
floating the photocatalyst in increasing the photocatalytic performance. provided the best conversion of 86.7%. Other contaminants were also
decomposed and/or removed by the photocatalytic treatment, CaCO3,
3.6. Reusability of photocatalyst COD, SO42−, BOD, and Cl− with the efficiency of 72.0, 56.0, 46.0, and
36.0, 28.5%, respectively. Also, the total hardness was diminished with
It is often found that photocatalytic reactions are completely sup- the efficiency of 53.4%. It was observed that floating the photocatalyst
pressed after a while, possibly because of backward reactions and caused to enhance the photocatalytic activity from 37% by pure ZnO to
poisoning effect that is a result of compounds chemically bounded on 86.7% by the floatable ZnO/Oak charcoal composite. The FTIR spectra
the catalyst surface [27]. Herein, to study the reusability, the ZnO/ of the ZnO/charcoal before and after three times recycling reveal the
charcoal particles were filtered and washed by deionized water three poisoning of the photocatalyst because of accumulation of the reactants
times. Then, they were poured into an aqueous HCl solution (pH 6) on the photocatalyst surface.
under stirring for 30 min to remove the likely deposited CaCO3 particles
on the photocatalyst surface based on Eq. (3). The likely SO42+ ions are Declaration of Competing Interest
also expected to be removed. At the next step, the photocatalyst par-
ticles were washed by an aqueous NaOH solution (pH 10.5) to remove The authors declare that they have no known competing financial
the remained HCl/Cl− based on Eq. (4). The surface remained NH4+ interests or personal relationships that could have appeared to influ-
can be converted to NH3 in basic media according to Eq. (5). Since NH3 ence the work reported in this paper.
can be easily removed from photocatalyst surface by steam regenera-
tion, the photocatalyst particles were exposed to air steam by an air Acknowledgement
diffuser in a Teflon vessel for 30 min (flow rate 2 L/min, temperature
120 °C). Finally, the photocatalyst particles were dried at 80 °C for 24 h We thank the Kermanshah Petrochemical Co. Iran, for providing the
followed by calcination at 300 °C for 1 h under nitrogen atmosphere. real ammonia wastewater and for assistance in some experimental
analyses.
CaCO3 (s) + 2 HCl CaCl2 (aq) + H2 O (l) + CO2 (g) (3)
References
−1 −
HCl (aq)/Cl + NaOH (aq) → NaCl (aq) + H2O (l)/OH (4)
[1] R.N. Nordin, L.W. Pommen, C.L. Meays, Water Quality Guidelines for Nitrogen
(Nitrate, Nitrite, and Ammonia), Accessed September 2009 https://www2.gov.bc.
NH+4 + OH NH3 + H2 O (5) ca/assets/gov/environment/air-land-water/water/waterquality/wqgs-wqos/
approved-wqgs/nitrogen-overview.pdf.
Fig. 8 shows the results of three times recycling the ZnO/charcoal [2] M.S. Vohra, S.M. Selimuzzaman, M.S. Al-Suwaiyan, NH4+-NH3 removal from si-
photocatalyst for the photocatalytic treatment of industrial ammonia mulated wastewater using UV-TiO2 photocatalysis: effect of co-pollutants and pH,
wastewater. As can be seen, the photocatalytic performance was di- Environ. Technol. 31 (6) (2010) 641–654.
[3] X. Zuo, J. Hu, M. Chen, The role and fate of inorganic nitrogen species during UVA/
minished obviously in second and third runs. Fig. 9 indicates the XRD TiO₂ disinfection, Water Res. 80 (2015) 12–19.
and FTIR patterns of the fresh and three times recycled ZnO/charcoal. [4] Y. Shavisi, S. Sharifnia, M. Zendehzaban, M.L. Mirghavami, S. Kakehazar,
From Fig. 9(a), it is obvious that the phase and structure of the used Application of solar light for degradation of ammonia in petrochemical wastewater
by a floating TiO2/LECA photocatalyst, J. Ind. Eng. Chem. 20 (5) (2014)
samples remained the same. However, the intensity of ZnO’s peaks
2806–2813.
decreased slightly due to catalyst losing. FTIR gives us more informa- [5] Y. Yang, Q. Ding, D. Wen, M. Yang, Y. Wang, N. Liu, X. Zhang, Bromate removal
tion on the possible changes in the surface chemical properties of from aqueous solution with novel flower-like Mg-Al-layered double hydroxides,
photocatalysts. As can be seen from Fig. 9(b), IR spectrum of the fresh Environ. Sci. Pollut. R. 25 (27) (2018) 27503–27513.
[6] N. Abdullah, N. Yusof, W.J. Lau, J. Jaafar, A.F. Ismail, Recent trends of heavy metal
ZnO/charcoals includes six main characteristic bands at 480, 620, 749, removal from water/wastewater by membrane technologies, J. Ind. Eng. Chem. 76
862, 1622 and 3410 cm−1. The band located at 480 cm−1 indicates the (25) (2019) 17–38.
stretching metal-oxide bond in ZnO structure. The bands at 620, 749 [7] K. Sivagami, K.P. Sakthivel, I.M. Nambi, Advanced oxidation processes for the
treatment of tannery wastewater, J. Environ. Chem. Eng. 6 (3) (2018) 3656–3663.
and 862 cm−1 are attributed to an aromatic network produced upon [8] D.B. Miklos, C. Remy, M. Jekelc, K.G. Linden, J.E. Drewes, U. Hübner, Evaluation of
carbonization process [48]. The bands located at 1622 and 3410 cm−1 advanced oxidation processes for water and wastewater treatment – a critical re-
are ascribed to the bending and stretching surface hydroxide groups. view, Water Res. 139 (2018) 118–131.
[9] A. Gholami Akerdi, S.H. Bahrami, Application of heterogeneous nano-semi-
Although, stretching vibration of the non-saturated C]C bonds may conductors for photocatalytic advanced oxidation of organic compounds: a review,
also the origin of absorption band at 1622 cm−1 [49]. The IR spectrum J. Environ. Chem. Eng. 7 (5) (2019) 103283.
of the used sample exhibits additional peaks at 781, 787 and [10] L. Kinidi, I.A. Wei Tan, N.B. Abdul Wahab, K.F. Bin Tamrin, C.N. Hipolito,
S.F. Salleh, Recent development in ammonia stripping process for industrial was-
1407 cm−1, the two formers may be assigned to the vibration modes of tewater treatment, Int. J. Chem. Eng. 2018 (2018).
NO2− and NO3− surface bonds [41] and/or chemisorbed Cl− on the [11] X. Quan, Z. Cheng, F. Xu, F. Qiu, L. Dai, Y. Yan, Structural optimization of the
charcoal (stretching C-Cl bond) [50] and the latter can be ascribed to porous section in a water-sparged aerocyclone reactor to enhance the air-stripping
efficiency of ammonia, J. Environ. Chem. Eng. 2 (2) (2014) 1199–1206.
the presence of calcite deposited on the photocatalyst surface [51].
[12] M.H. Yuan, Y.H. Chen, J.Y. Tsai, C.Y. Chang, Removal of ammonia from wastewater
Therefore, it can be said that the declined catalytic performance after by air stripping process in laboratory and pilot scales using a rotating packed bed at
three times recycling is due to the loss of catalyst as well as the poi- ambient temperature, J. Taiwan Inst. Chem. E 60 (2016) 488–495.
[13] T.C. Jorgensen, L.R. Weatherley, Ammonia removal from wastewater by ion ex-
soning effect caused by accumulated of the reactants on catalyst sur-
change in the presence of organic contaminants, Water Res. 37 (8) (2003)
face. 1723–1728.
[14] R. Malekian, J. Abedi-Koupai, S.S. Eslamian, S.F. Mousavi, K.C. Abbaspour,
M. Afyuni, Ion-exchange process for ammonium removal and release using natural
4. Conclusion Iranian zeolite, Appl. Clay Sci. 51 (3) (2011) 323–329.
[15] X. Yang, C. Shang, J.C. Huang, DBP formation in breakpoint chlorination of was-
This study deals with the photocatalytic decontamination of tewater, Water Res. 39 (19) (2005) 4755–4767.

10
N. Tafreshi, et al. Journal of Environmental Chemical Engineering 7 (2019) 103397

[16] J. Jin, M.G. El-Din, J.R. Bolton, Assessment of the UV/chlorine process as an ad- [36] L.C. Machado, C.B. Torchia, R.M. Lago, Floating photocatalysts based on TiO2
vanced oxidation process, Water Res. 45 (4) (2011) 1890–1896. supported on high surface area exfoliated vermiculite for water decontamination,
[17] L.L. Albornoz, L. Marder, T. Benvenuti, A.M. Bernardes, Electrodialysis applied to Catal. Commun. 7 (8) (2006) 538–541.
the treatment of an university sewage for water recovery, J. Environ. Chem. Eng. 7 [37] N.B. Jackson, C.M. Wang, Z. Luo, J. Schwitzgebel, J.G. Ekerdt, J.R. Brock, A. Heller,
(2) (2019) p.102982. Attachment of TiO2 powders to hollow glass microbeads: activity of the
[18] D. Ippersiel, M. Mondor, F. Lamarche, F. Tremblay, J. Dubreuil, L. Masse, Nitrogen TiO2‐coated beads in the photoassisted oxidation of ethanol to acetaldehyde, J.
potential recovery and concentration of ammonia from swine manure using elec- Electrochem. Soc. 138 (12) (1991) 3660–3664.
trodialysis coupled with air stripping, J. Environ. Manage. 95 (2012) S165–S169. [38] L. Ni, Y. Li, C. Zhang, L. Li, W. Zhang, D. Wang, Novel floating photocatalysts based
[19] G. Ruiz, D. Jeison, O. Rubilar, G. Ciudad, R. Chamy, Nitrification–denitrification via on polyurethane composite foams modified with silver/titanium dioxide/graphene
nitrite accumulation for nitrogen removal from wastewaters, Bioresour. Technol. 97 ternary nanoparticles for the visible‐light‐mediated remediation of diesel‐polluted
(2) (2006) 330–335. surface water, J. Appl. Polym. Sci. 133 (19) (2016).
[20] F. Jaramillo, M. Orchard, C. Munoz, M. Zamorano, C. Antileo, Advanced strategies [39] S. Singh, P.K. Singh, H. Mahalingam, Novel floating Ag+-doped TiO2/polystyrene
to improve nitrification process in sequencing batch reactors – a review, J. Environ. photocatalysts for the treatment of dye wastewater, Ind. Eng. Chem. Res. 53 (42)
Manage. 218 (2018) 154–164. (2014) 16332–16340.
[21] T.P. O’Farell, F.P. Frauson, A.F. Cassel, D.F. Bishop, Nitrogen removal by ammonia [40] I. Altın, M. Sökmen, Preparation of TiO2-polystyrene photocatalyst from waste
stripping, J. Water Pollut. Control Fed. 44 (8) (1972) 1527–1535. material and its usability for removal of various pollutants, Appl. Catal. B 144
[22] J.C. Campos, A.P. Moura, L. Costa, F.V. Yokoyama, D.F. Arouja, M.C. Cammarota, (2014) 694–701.
Evaluation of pH, alkalinity and temperature during air stripping process for am- [41] N. Tafreshi, S. Sharifnia, S. Moradi Dehaghi, Box–Behnken experimental design for
monia removal from landfill leachate, J Environ. Sci. Health 48 (9) (2013) optimization of ammonia photocatalytic degradation by ZnO/Oak charcoal com-
1105–1113. posite, Process Saf. Environ. Prot. 106 (2017) 203–210.
[23] M. Ellersdorfer, The ion-exchange-loop stripping process: ammonium recovery from [42] T. Rakić, I. Kasagić-Vujanović, M. Jovanović, B. Jančić-Stojanović, D. Ivanović,
sludge liquor using NACl-treated clinoptilolite and simultaneous air stripping, Comparison of full factorial design, central composite design, and box-behnken
Water Sci. Technol. 77 (3) (2017) 695–705. design in chromatographic method development for the determination of flucona-
[24] F.M. Ferraz, J. Povinelli, E.M. Veira, Ammonia removal from landfill leachate by air zole and its impurities, Anal. Lett. 47 (8) (2014) 1334–1347.
stripping and absorption, Environ. Technol. 34 (15) (2013) 2317–2326. [43] S.C. Ferreira, R.E. Bruns, H.S. Ferreira, G.D. Matos, J.M. David, G.C. Brandao,
[25] X. Zhang, Y. Yang, W. Huang, Y. Yang, Y. Wang, C. He, N. Liu, M. Wu, L. Tang, g- E.P. da Silva, L.A. Portugal, P.S. Dos Reis, A.S. Souza, W.N.L. Dos Santos, Box-
C3N4/UiO-66 nanohybrids with enhanced photocatalytic activities for the oxidation Behnken design: an alternative for the optimization of analytical methods, Anal.
of dye under visible light irradiation, Mater. Res. Bull. 99 (2018) 349–358. Chim. Acta 597 (2) (2007) 179–186.
[26] N. Liu, W. Huang, M. Tang, C. Yin, B. Gao, Z. Li, L. Tang, J. Lei, L. Cui, X. Zhang, In- [44] S. Mobini, F. Meshkani, M. Rezaei, Synthesis and characterization of nanocrystal-
situ fabrication of needle-shaped MIL-53 (Fe) with 1T-MoS2 and study on its en- line copper–chromium catalyst and its application in the oxidation of carbon
hanced photocatalytic mechanism of ibuprofen, Chem. Eng. J. 359 (2019) 254–264. monoxide, Process Saf. Environ. Prot. 107 (2017) 181–189.
[27] Y. Shaveisi, S. Sharifnia, E. Karamian, Application of mixture experimental design [45] S. Mobini, F. Meshkani, M. Rezaei, Surfactant-assisted hydrothermal synthesis of
for photocatalytic ammonia degradation by sunlight‐driven WO3‐Ag3PO4‐ZnO CuCr2O4 spinel catalyst and its application in CO oxidation process, J. Environ.
ternary photocatalysts, Int. J. Energy Res. 43 (9) (2019) 4879–4897. Chem. Eng. 5 (5) (2017) 4906–4916.
[28] H. Koohestani, S.K. Sadrnezhaad, Photocatalytic activity of immobilized geometries [46] X. Zhang, X. Zhang, L. Song, F. Hou, Y. Yang, Y. Wang, N. Liu, Enhanced catalytic
of TiO2, J. Mater. Eng. Perform. 24 (7) (2015) 2757–2763. performance for CO oxidation and preferential CO oxidation over CuO/CeO2 cat-
[29] Y. Shavisi, S. Sharifnia, Z. Mohamadi, Solar-light-harvesting degradation of aqueous alysts synthesized from metal organic framework: effects of preparation methods,
ammonia by CuO/ZnO immobilized on pottery plate: linear kinetic modeling for Int. J. Hydrogen Energy 43 (39) (2018) 18279–18288.
adsorption and photocatalysis process, J. Environ. Chem. Eng. 4 (3) (2016) [47] X. Zhang, X. Lv, X. Shi, Y. Yang, Y. Yang, Enhanced hydrophobic UiO-66 (University
2736–2744. of Oslo 66) metal-organic framework with high capacity and selectivity for toluene
[30] E. Karamian, S. Sharifnia, Enhanced visible light photocatalytic activity of BiFeO3- capture from high humid air, J. Colloid Interf. Sci. 539 (2019) 152–160.
ZnO pn heterojunction for CO2 reduction, Mater. Sci. Eng. B 238 (2018) 142–148. [48] Z.U.O. Song-lin, G. Shang-yu, Y. Xi-gen, X. Bo-sen, Carbonization mechanism of
[31] X. Zhang, Y. Wang, F. Hou, H. Li, Y. Yang, X. Zhang, Y. Yang, Y. Wang, Effects of Ag bamboo (phyllostachys) by means of Fourier Transform Infrared and elemental
loading on structural and photocatalytic properties of flower-like ZnO micro- analysis, J. Forestry Res. 14 (1) (2003) 75–79.
spheres, Appl. Surf. Sci. 391 (2017) 476–483. [49] C.D. Zappielo, D.M. Nanicuacua, W.N. dos Santos, D.L. da Silva, L.H. Dall’Antônia,
[32] Y. Yang, H. Li, F. Hou, J. Hu, X. Zhang, Y. Wang, Facile synthesis of ZnO/Ag na- F.M.D. Oliveira, D.N. Clausen, C.R. Tarley, Solid phase extraction to on-line pre-
nocomposites with enhanced photocatalytic properties under visible light, Mater. concentrate trace cadmium using chemically modified nano-carbon black with 3-
Lett. 180 (2016) 97–100. mercaptopropyltrimethoxysilane, J. Brazil. Chem. Soc. 27 (10) (2016) 1715–1726.
[33] M. Długosz, J. Waś, K. Szczubiałka, M. Nowakowska, TiO2-coated EP as a floating [50] L. Wang, B. Wu, W. Li, S. Wang, Z. Li, M. Li, D. Pan, M. Wu, Amphiphilic graphene
photocatalyst for water purification, J. Mater. Chem. A 2 (19) (2014) 6931–6938. quantum dots as self‐targeted fluorescence probes for cell nucleus imaging, Adv.
[34] M. Zendehzaban, S. Sharifnia, S.N. Hosseini, Photocatalytic degradation of am- Biosyst. 2 (8) (2018) p 1700191.
monia by light expanded clay aggregate (LECA)-coating of TiO2 nanoparticles, [51] G.B. Cai, S.F. Chen, L. Liu, J. Jiang, H.B. Yao, A.W. Xu, S.H. Yu, 1, 3-Diamino-2-
Korean J. Chem. Eng. 30 (3) (2013) 574–579. hydroxypropane-N, N, N′, N′-tetraacetic acid stabilized amorphous calcium carbo-
[35] Y. Shavisi, S. Sharifnia, S.N. Hosseini, M.A. Khadivi, Application of TiO2/perlite nate: nucleation, transformation and crystal growth, CrystEngComm 12 (1) (2010)
photocatalysis for degradation of ammonia in wastewater, J. Ind. Eng. Chem. 20 (1) 234–241.
(2014) 278–283.

11

You might also like