Physical Chemistry
p-ISSN: 2167-7042 e-ISSN: 2167-7069
2017; 7(2): 42-53
doi:10.5923/j.pc.20170702.03

Taha M. Elmorsi
Chemistry Department, Faculty of Science, Al-Azhar University, Cairo, Egypt
Correspondence to: Taha M. Elmorsi, Chemistry Department, Faculty of Science, Al-Azhar University, Cairo, Egypt.
| Email: | ![]() |
Copyright © 2017 Scientific & Academic Publishing. All Rights Reserved.
This work is licensed under the Creative Commons Attribution International License (CC BY).
http://creativecommons.org/licenses/by/4.0/

Photocatalytic properties of photocatalysts depend generally on techniques employed for preparing nanostructures and doping processes. Nanostructured zinc oxides (ZnO) are promising photocatalysts due to their high stability, high surface activity, and low costs. This study highlights a simple method of a co-precipitation for synthesizing ZnO and potassium doped ZnO (K-ZnO). In this research, surface functionalization with high quality explored by multiple surface analytical methods such as crystal structural, morphology analysis, inductively coupled plasma-atomic emissions spectroscopy (ICP-AES) and scanning electron microscopy (SEM) attached with energy dispersive x-ray (SEM/EDX). The approach is well suited to produce crystalline K-ZnO, which possesses a shape of nanorods (35.2 to 48.9 nm) and acts as an efficient visible light responsive photocatalyst. The results indicated that the K-ZnO is more responsive to the visible light than pure ZnO. The degradation efficiency of 2-naphthol in aqueous solution in the presence of K-ZnO and sunlight was found to be 95% in 70 min at pH 10.5. In contrast, negligible degradation of 2-naphthol occurred in the absence of K-ZnO. Mechanism and type of photoreaction were identified by employing hydroxyl radical scavengers (e.g. tert-butyl alcohol and Cu2+/Ag+ ion) and electron
scavengers (e.g. Ag+ ions and molecular oxygen). The kinetic analysis of the photodegradation process using a Langmuir-Hinshelwood model suggests that the reaction has occurred on the surface of the photocatalyst. Also, the reaction is influenced by hydroxyl radical scavengers supporting a photooxidation pathway, whereas electron scavengers showed minor effect. Mineralization of 2-naphthol using K-ZnO in presence of sunlight was confirmed by ultra-performance liquid chromatography equipped with mass spectrometry (UPLC/MS).
Keywords: Visible-Light photocatalysis, ZnO, Doping process, Scavengers, Langmuir-Hinshelwood, 2-naphthol
Cite this paper: Taha M. Elmorsi, Toward Visible-Light Responsive Photocatalysts: Nano-Potassium Doping Zinc Oxide (K-ZnO) for Degradation of 2-Naphthol, Physical Chemistry, Vol. 7 No. 2, 2017, pp. 42-53. doi: 10.5923/j.pc.20170702.03.
and valence band holes
(equation 1). The active species
may react with
and
to produce hydroxyl radicals
and super oxide anion radical
respectively (equations 2-3). Photooxidation and photoreduction reactions of pollutants may occur directly and/or indirectly (equation 4-5) to produce different products (P) [16, 17].![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
radicals or via free solvated electron [13, 18]. Although, the applicability of experimental data to the Langmuir-Hinshelwood kinetic model would support a surface reaction, further evidences are required [13, 19]. Recently, researchers focused on improving the photocatalytic properties of the photocatalysts using different techniques for preparing nanostructures and doping processes [15]. Nanostructured ZnO has been synthesized by different techniques such as hydrothermal method, aerosol, micro-emulsion, ultrasonic, sol–gel method, evaporation of solution and suspensions, solid state reaction, wet chemical synthesis and other techniques [20-22]. However, among these methods, co-precipitation route appeared to be a simple route, low cost and provide large scale production of effective nanoparticle ZnO [22, 23]. It is known that synthesis of ZnO nanoparticles in different shapes with different particle sizes, would lead to increase its surface area which may improve its photocatalytic properties [24, 25]. On the other hand, incorporating dopant ions would generate lattice defects and lead to change in the band gap energy. Furthermore, lowering the band gap of ZnO is an efficient method for enhancing the visible light response of ZnO. Initiating photocatalytic reactions with visible light energy is the main need to overcome the cost and to use the renewable energy. The objective of this research was to use a simple method of co-precipitation for synthesizing nanoparticles of pure ZnO and potassium doped ZnO (K-ZnO) as a visible light responsive photocatalyst. K-ZnO nanoparticles will be used for photocatalytic degradation of 2-naphthol (an organic pollutant) using direct sun light. The optimum conditions for photocatalytic degradation will be determined and different
and
scavengers will be used to elucidate the type of photocatalytic reaction.
(mg/g)), and at equilibrium (
(mg/g)) calculated as in equations 6 and 7.![]() | (6) |
![]() | (7) |
and
are concentration (mg/L) of 2-naphthol before starting the experiment, concentration during the experiment and equilibrium concentrations. Solution volume is V(L) and the photocatalyst mass is W(g). The removal percentage of 2-naphthol by photocatalytic degradation can be calculated from equation 8:![]() | (8) |
![]() | (9) |
![]() | (10) |
![]() | (11) |
![]() | (12) |
and
are the reaction rate constants of first-order and second-order reactions respectively.![]() | Figure 1. X-ray diffraction patterns with (hkl) values of as-prepared K-ZnO nanorods compared to the standard (JCPDS:36-1451) of pure ZnO |
![]() | Figure 2. SEM image (a) and EDX spectrum (b) of as prepared K-ZnO sample |
|
![]() | Figure 3. TEM image of K-ZnO nanorods prepared by co-precipitation |
![]() | Figure 4. FTIR spectra of K-ZnO nanorods and pure ZnO samples |
![]() | (12) |
are the absorption coefficient and the band tailing parameter (energy-independent constant) respectively,
is the optical energy gap,
is the power factor of the transition mode (in this study n =2),
(is the photon energy) where
is the plank’s constant and
is the frequency of the radiation. It is noted that the term
in equation (12) has different values based on the nature of the materials (amorphous or crystalline) and the type of the transition (allowed, not allowed and forbidden) as 1/2, 2, 3/2 and 3. In this study the best fit of our experimental data was obtained with n =2. The values of
of K-ZnO was determined by plotting
versus
, (as shown in Figure 5b) based on the rearrangement of Tauc’s model equation 12 becomes:![]() | (13) |
) axis at
. The obtained values were 2.94 and 3.21 eV for K-ZnO and pure ZnO respectively. This result indicated that the K-ZnO is more responsive to the visible light than pure ZnO. Also, confirmed the effectiveness of K-ZnO for the photodegradation of 2-naphthol in sunlight.![]() | Figure 5. UV-vis spectra of (a) the as-prepared K-ZnO and pure ZnO and (b) the estimated band gap energies |
and valence band holes
which are capable of initiating photoreduction and photooxidation of 2-naphthol, respectively.The change in the concentrations of 2-naphthol
with illumination time, indicated that about 95% disappearance occurred in 70 min of irradiation. This result can be explained in terms of presence of K in K doped ZnO nanoparticles in addition to its small particles size (35.2 nm to 48.9 nm) as determined by TEM analysis. Presence of K in K-ZnO led to increasing the response of K-ZnO to the visible light with enough energy to promote the photo redox species that are capable of photodegradation of 2-naphthol [28]. Furthermore, nanoparticles of K-ZnO would have large surface area which possess more available surface sits for photodegradation process [32]. However, in presence of pure ZnO under similar conditions the photocatalytic degradation of 2-naphthol was very small with only about 21% disappearance in same photoperiod. Furthermore, it was noted that both direct photolysis and dark adsorption led to slight remove to 2-naphthol. Hence removal of 2-naphthol due to direct photolysis and dark adsorption was only 12% and 16% respectively over similar conditions.
radicals which are the key for the degradation of 2-naphthol. Similar results were reported for different studies [32, 33]. ![]() | Figure 7. Effect of pH value on photocatalytic degradation of 2-naphthol, [2-naphthol]0 = 12 mg/L, dose of K-ZnO = 0.5 g/L, T = 37 °C and pH = 3.5 to 10.5 under Sunlight |
|
radicals to attack high number of 2-naphthol molecules and increase the degradation rate constant as shown in Table 3. Similar results were reported previously for the degradation of congo red by ZnO [34]. In subsequent experiments 0.5 g/L of K-ZnO was used to avoid using more dose of catalyst.![]() | Figure 8. Effect of K-ZnO dose on photocatalytic degradation of 2-naphthol, [2-naphthol]0 = 12 mg/L, dose of K-ZnO = 0.5 g/L, T = 37°C and pH = 3.5 to 10.5 under Sunlight |
|
radicals [32]. Hence decreases the degradation rate of 2-naphthol. This results were in a good agreement with previous reports [34, 36]. To investigate the relationship between the initial rate of photodegradation and the initial concentration of the pollutant, Langmuir–Hinshelwood (L–H) equation was used [13, 32]. This model is describing the photocatalytic degradation process occurred onto the surface of the photocatalyst. According to the linear form of L-H (equation 14), the reciprocal initial rate is plotted versus the reciprocal initial concentration as shown in Figure (9b). ![]() | (14) |
, is initial rate of the photodegradation, k is the reaction rate constant (mol/l min),
is the Langmuir-Hinshelwood equilibrium adsorption coefficient (l/mol) and
is the initial concentration of 2-naphthol. The experimental data showed a good fit with L-H model. The values of the reaction rate constant (k) obtained as 0.681 (mol/l. min) and the adsorption coefficient of the reactant (KH-L) as 18350 (l/mol). The linearity of L-H model along with resendable value of K indicated that the photocatalytic degradation of 2-naphthol is taking place mainly on the surface of K-ZnO. It is better to note that in all subsequent experiments the initial concentration of 2-naphthol was used as 12 mg/L to investigate the effect of other parameters.
of K-ZnO at equilibrium are calculated according to equation 7.While the equilibrium concentrations
in the solution calculated from the calibration curve and the absorbance of 2-napthol at λmax of 228nm. The relation between
and
at constant temperature which is called the adsorption isotherm is shown in Figure 10a. The adsorption isotherm of this experiments was found to follow the Langmuir model (equation 15) with very high coefficient of determination
. ![]() | (15) |
is the maximum amount of 2-naphthol adsorbed on the surface of K-ZnO at a complete monolayer (mg/g) and
is the Langmuir adsorption constant (L/mg) which is related to the free energy of adsorption. A plot of
versus
for the adsorption of 2-naphthol (Figure 10b) gives a straight line of slope
and intercept
which are used to calculate Langmuir constants. The maximum adsorption capacity of 2-naphthol on K-ZnO
was found to be 3.29 x10-5 mol/g (4.74 mg/g) and the Langmuir constant
is 1496.70 (L/mol). As discussed previously, the value of adsorption constant calculated from the Langmuir–Hinshelwood model (KL-H) for the photocatalytic degradation is 18350 L/mol. It can be noted that the adsorption rate of 2-naphthol in the light is 12 times more than the adsorption in the dark. It is indicated that the irradiated light would affect the adsorption rate along with the electronic properties of K-ZnO [32, 37]. Several studies have shown similar results for different compounds [38, 39]. It can be concluded that the adsorption of 2- naphthol on the surface of K-ZnO would enhance the degradation rate by the adsorbed
radicals [13], However, the adsorption process is not only the key for the degradation of the pollutant, hence the adsorbed
radicals may diffuse into the bulk solution and continue the process [37]. ![]() | Figure 10. (a) Adsorption isotherm of 2-naphthol on K-ZnO surface and (b) Langmuir isotherm model, [2-naphthol]0 = 3.0 to 24 mg/L, dose of K-ZnO = 0.5 g/L, T = 37°C, pH = 10.5 and Time = 4.0 hr |
and
radical scavengers on the rate was mainly used to provide valuable information. Different scavengers were added to eliminate one route of degradation at a time.
scavengers. In this experiment, electron scavenger is useful to confirm that the photocatalytic degradation of 2-naphthol was an oxidative process. It is expected that electron scavengers would inhibit the rate of degradation of 2-naphthol if it was being reduced by reaction with
. While, the presence of the
scavengers could enhance the rate of photocatalytic degradation if 2-naphthol was being oxidized by
[13]. Ag+ ion, is known as an effective electron scavenger due to its reduction by
to metallic silver
according to equation (16):![]() | (16) |
radicals by a decrease in the degree of electron-hole recombination [13]. This result confirmed that the
radicals are the main oxidizing species for degrading 2-naphthol by photooxidation process. Furthermore, molecular oxygen is another
scavenger due to its conversion to the superoxide anion radical
[13]. Then
will eventually reproduce adsorbed oxygen
. The photocatalytic degradation of 2-naphthol was rapid in presence of the natural level of O2 (line namely no scavengers in Fig.11) in the solution (without purging more O2). While decreasing the level of dissolved oxygen by N2 purging in the solution for an hour, led to a significant decrease in the degradation rate by about 30%. It appears that reducing the scavenging effect of
enhances the degree of electron-hole recombination and decreases the available amount of
radicals. Decreasing the degradation rate of 2-naphthol due to the absence of more
radicals is further confirming for the photooxidation process. ![]() | Figure 11. Effect of electron scavengers on the photocatalytic degradation of 2-naphthol, [2-naphthol]0 = 12 mg/L, dose of K-ZnO = 0.5 g/L, T = 37°C and pH = 10.5 under Sunlight |
Radical Scavengers
radical scavengers. Tert-butyl alcohol (TBA) and Cu2+ in presence of Ag+ ions were used. The result in Figure 12 showed that the degradation rate of 2-naphthol, was significantly decreased in the presence of TBA. Hence TBA is weakly adsorbed onto the surface of a photocatalyst [40], it is mainly reacted with free
radicals
which available for photooxidation. Presence of TBA inhibited the degradation of 2-naphthol as shown in Figure 12. Furthermore, the photodegradation of 2-napthol was strongly inhibited due to the addition of 1.5 mM of Cu2+ ion in presence of 1.5mM Ag+ ions lending further support for an oxidative degradation. Presence of Ag+ ions in the solution reduced Cu2+ ions to unstable Cu+. Consequently, the reactive Cu+ ions formed [13] compete with 2-naphthol for the oxidants,
according to equation 17.![]() | (17) |
radicals either free or adsorbed on the surface of K-ZnO. It is worthy to note that the adsorption experiments discussed previously, further confirmed the photooxidation process via the
radicals. ![]() | Figure 12. Effect of radical Scavengers on the photocatalytic degradation of 2-naphthol, [2-naphthol]0 = 12 mg/L, dose of K-ZnO = 0.5 g/L, T = 37°C and pH = 10.5 under Sunlight |
![]() | Figure 13. (a) UPLC chromatograms at 228 nm (and TIC obtained at 8.93 min is 2-naphthol at different time of photocatalytic degradation; 0.0 min, 30 min and 70 min, (b) mass spectra of TIC at 8.93 min (for 2-naphthol) m/z = 144.95 |
and
supported a photooxidation reaction. UPLC/MS analysis confirmed the mineralization of 2-naphthol using K-ZnO in presence of sunlight. | [1] | Pirilä, M., et al., Photocatalytic degradation of organic pollutants in wastewater. Topics in Catalysis, 2015. 58(14-17): p. 1085-1099. |
| [2] | Qourzal, S., et al., Photodegradation of 2-naphthol in water by artificial light illumination using TiO2 photocatalyst: identification of intermediates and the reaction pathway. Applied Catalysis A: General, 2008. 334(1): p. 386-393. |
| [3] | González, L.T., et al., Comparative Photocatalytic Performance on the Degradation of 2-Naphthol Under Simulated Solar Light Using α-Bi4V2O11 Synthesized by Solid-State and Co-precipitation Methods. Water, Air, & Soil Pollution, 2017. 228(2): p. 75. |
| [4] | Luo, H., et al., Electrochemical degradation of phenol by in situ electro-generated and electro-activated hydrogen peroxide using an improved gas diffusion cathode. Electrochimica Acta, 2015. 186: p. 486-493. |
| [5] | Benhebal, H., et al., Photocatalytic degradation of phenol and benzoic acid using zinc oxide powders prepared by the sol–gel process. Alexandria Engineering Journal, 2013. 52(3): p. 517-523. |
| [6] | Delval, F., G. Crini, and J. Vebrel, Removal of organic pollutants from aqueous solutions by adsorbents prepared from an agroalimentary by-product. Bioresource technology, 2006. 97(16): p. 2173-2181. |
| [7] | Liu, C., et al., Thermal conversion of lignin to phenols: Relevance between chemical structure and pyrolysis behaviors. Fuel, 2016. 182: p. 864-870. |
| [8] | Zhou, W., et al., Phenol degradation by Sulfobacillus acidophilus TPY via the meta-pathway. Microbiological Research, 2016. 190: p. 37-45. |
| [9] | Zhou, M., et al., Investigation on the preparation and properties of monodispersed Al2O3 ZrO2 nanopowder via Co-precipitation method. Journal of Alloys and Compounds, 2016. 678: p. 337-342. |
| [10] | Wang, T., et al., Research on degradation product and reaction kinetics of membrane electro-bioreactor (MEBR) with catalytic electrodes for high concentration phenol wastewater treatment. Chemosphere, 2016. 155: p. 94-99. |
| [11] | Aslam, M., et al., Evaluation of sunlight induced structural changes and their effect on the photocatalytic activity of V2O5 for the degradation of phenols. Journal of hazardous materials, 2015. 286: p. 127-135. |
| [12] | Quiñones, D., et al., Boron doped TiO2 catalysts for photocatalytic ozonation of aqueous mixtures of common pesticides: Diuron, o-phenylphenol, MCPA and terbuthylazine. Applied Catalysis B: Environmental, 2015. 178: p. 74-81. |
| [13] | El-Morsi, T.M., et al., Photocatalytic degradation of 1, 10-dichlorodecane in aqueous suspensions of TiO2: A reaction of adsorbed chlorinated alkane with surface hydroxyl radicals. Environmental science & technology, 2000. 34(6): p. 1018-1022. |
| [14] | Ravelli, D., et al., Photocatalysis. A multi-faceted concept for green chemistry. Chemical Society Reviews, 2009. 38(7): p. 1999-2011. |
| [15] | Raj, R.B., et al., Effect of potassium on structural, photocatalytic and antibacterial activities of ZnO nanoparticles. Advances in Natural Sciences: Nanoscience and Nanotechnology, 2016. 7(4): p. 045008. |
| [16] | Pelizzetti, E. and C. Minero, Mechanism of the photo-oxidative degradation of organic pollutants over TiO2 particles. Electrochimica acta, 1993. 38(1): p. 47-55. |
| [17] | Calza, P., C. Minero, and E. Pelizzetti, Photocatalytically assisted hydrolysis of chlorinated methanes under anaerobic conditions. Environmental science & technology, 1997. 31(8): p. 2198-2203. |
| [18] | Kerzig, C. and M. Goez, Generating hydrated electrons through photoredox catalysis with 9-anthrolate. Physical Chemistry Chemical Physics, 2015. 17(21): p. 13829-13836. |
| [19] | Hoffmann, M.R., et al., Environmental applications of semiconductor photocatalysis. Chemical reviews, 1995. 95(1): p. 69-96. |
| [20] | Kołodziejczak-Radzimska, A. and T. Jesionowski, Zinc oxide—from synthesis to application: a review. Materials, 2014. 7(4): p. 2833-2881. |
| [21] | Kale, R.B., Morphological evolution of hydrothermally derived ZnO nano and microstructures. Optik-International Journal for Light and Electron Optics, 2016. 127(11): p. 4621-4624. |
| [22] | Sadraei, R., A Simple Method for Preparation of Nano-sized ZnO. Research & Reviews: Journal of Chemistry, e-ISSN, 2016: p. 2319-9849. |
| [23] | Chen, C., P. Liu, and C. Lu, Synthesis and characterization of nano-sized ZnO powders by direct precipitation method. Chemical Engineering Journal, 2008. 144(3): p. 509-513. |
| [24] | Duo, S., et al., A facile salicylic acid assisted hydrothermal synthesis of different flower-like ZnO hierarchical architectures with optical and concentration-dependent photocatalytic properties. Materials Characterization, 2016. 114: p. 185-196. |
| [25] | Xiao, Q., et al., Sonochemical synthesis of ZnO nanosheet. Journal of Alloys and Compounds, 2008. 459(1): p. L18-L22. |
| [26] | Chen, X., et al., Self-assembly of ZnO nanoparticles into hollow microspheres via a facile solvothermal route and their application as gas sensor. CrystEngComm, 2013. 15(36): p. 7243-7249. |
| [27] | Hassanien, A. and A.A. Akl, Crystal imperfections and Mott parameters of sprayed nanostructure IrO2 thin films. Physica B: Condensed Matter, 2015. 473: p. 11-19. |
| [28] | Bharathi, V., et al., Optical, structural, enhanced local vibrational and fluorescence properties in K-doped ZnO nanostructures. Applied Physics A, 2014. 116(1): p. 395-401. |
| [29] | Djaja, N.F., D.A. Montja, and R. Saleh, The effect of Co incorporation into ZnO nanoparticles. 2013. |
| [30] | Ansari, S.A. and M.H. Cho, highly visible light responsive, narrow band gap TiO2 nanoparticles modified by elemental red phosphorus for photocatalysis and photoelectrochemical applications. Scientific reports, 2016. 6. |
| [31] | Tumuluri, A., K.L. Naidu, and K.J. Raju, Band gap determination using Tauc’s plot for LiNbO3 thin films. Chem Tech, 2014. 6(6): p. 3353-3356. |
| [32] | Laid, n., et al., study of the photocatalytic degradation of a cationic dye in aqueous solution by different types of catalysts. Sciences & technologie a, 2014(39): p. 75-82. |
| [33] | Konstantinou, I.K. and T.A. Albanis, TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: kinetic and mechanistic investigations: a review. Applied Catalysis B: Environmental, 2004. 49(1): p. 1-14. |
| [34] | Zuas, O., H. Budiman, and N. Hamim, Synthesis of ZnO nanoparticles for microwave induced rapid catalytic decolorization of congo red dye. Adv. Mater. Lett, 2013. 4: p. 662-667. |
| [35] | Teh, C.M. and A.R. Mohamed, Roles of titanium dioxide and ion-doped titanium dioxide on photocatalytic degradation of organic pollutants (phenolic compounds and dyes) in aqueous solutions: a review. Journal of Alloys and Compounds, 2011. 509(5): p. 1648-1660. |
| [36] | Bi, X.-y., et al., Treatment of phenol wastewater by microwave-induced ClO2-CuOx/Al2O3 catalytic oxidation process. Journal of Environmental Sciences, 2007. 19(12): p. 1510-1515. |
| [37] | Qourzal, S., et al., Photocatalytic degradation and adsorption of 2-naphthol on suspended TiO2 surface in a dynamic reactor. Journal of Colloid and Interface Science, 2005. 286(2): p. 621-626. |
| [38] | Huang, H.-H., D.-H. Tseng, and L.-C. Juang, Heterogeneous photocatalytic degradation of monochlorobenzene in water. Journal of Hazardous Materials, 2008. 156(1): p. 186-193. |
| [39] | Mills, A. and S. Morris, Photomineralization of 4-chlorophenol sensitized by titanium dioxide: a study of the initial kinetics of carbon dioxide photogeneration. Journal of Photochemistry and Photobiology A: Chemistry, 1993. 71(1): p. 75-83. |
| [40] | Ramakrishnan, A., et al., Enhanced performance of surface-modified TiO2 photocatalysts prepared via a visible-light photosynthetic route. Chemical Communications, 2012. 48(68): p. 8556-8558. |
| [41] | Martinez-de La Cruz, A., et al., Photoassisted degradation of rhodamine B by nanoparticles of α-Bi2Mo3O12 prepared by an amorphous complex precursor. Materials Chemistry and Physics, 2008. 112(2): p. 679-685. |
| [42] | Mart nez-de la Cruz, A. and S. Obreg n Alfaro, Synthesis and characterization of nanoparticles of?-Bi2Mo3O12 prepared by co-precipitation method: Langmuir adsorption parameters and photocatalytic properties with rhodamine B. Solid State Sciences, 2009. 11(4): p. 829-835. |
| [43] | Yuan, R., et al., Effects of chloride ion on degradation of Acid Orange 7 by sulfate radical-based advanced oxidation process: implications for formation of chlorinated aromatic compounds. Journal of hazardous materials, 2011. 196: p. 173-179. |