Growth Mechanism of Sea Urchin ZnO Nanostructures in Aqueous

Aug 14, 2019 - This work reports the development of a fast and simple route for the synthesis of ZnO sea urchin (SU) nanostructures by the formation a...
0 downloads 0 Views 8MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega XXXX, XXX, XXX−XXX

http://pubs.acs.org/journal/acsodf

Growth Mechanism of Sea Urchin ZnO Nanostructures in Aqueous Solutions and Their Photocatalytic Activity for the Degradation of Organic Dyes Hiran D. Kiriarachchi, Khaled M. Abouzeid, Longli Bo,† and M. Samy El-Shall* Department of Chemistry, Virginia Commonwealth University, Richmond, Virginia 23284-2006, United States

Downloaded via 37.44.252.154 on August 15, 2019 at 16:14:05 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: This work reports the development of a fast and simple route for the synthesis of ZnO sea urchin (SU) nanostructures by the formation and assembly of ZnO nanorods under favorable growth conditions in an aqueous solution. The thermal treatment of a basic zinc acetate solution in ethanol results in the formation of aggregated seed clusters consisting of small ZnO nanorods, which were then grown in a precursor solution containing Zn(NO3)2·6H2O and hexamethylenetetramine to assemble the SU structures from the anisotropic ZnO nanorods on the surface of the seed clusters. Each ZnO nanoparticle in the aggregated seed clusters grew sequentially into a ZnO nanorod, and the nanorods were concentric to the core of the clusters yielding the unique SU-like shape. In the presence of a capping agent such as cetyl trimethyl ammonium bromide (CTAB), the aggregated seed clusters were not formed, and the growth of the CTAB-capped ZnO nanorods resulted in separated rods with average aspect ratios of ∼10. The SU ZnO nanostructures exhibit a hexagonal wurtzite crystal structure and higher specific surface area (26.9 m2/g) than the CTAB-capped nanorods (17.7 m2/g). The SU ZnO nanostructures show superior photocatalytic efficiency for the degradation of three common organic dyes compared to the ZnO nanorods. The removal efficiencies of indigo carmine, methylene blue, and rhodamine B by the SU nanostructures were 99, 86, and 96%, respectively, after 1 h of UV irradiation. Therefore, the ZnO SU structures have the potential to be a versatile photocatalyst for the photodegradation of organic dyes in industrial wastewater.

1. INTRODUCTION In recent years, 3-dimensional (3D) semiconductor metal oxide nanostructures have gained considerable attention due to their wide range of applications in numerous fields.1−4 However, very few semiconductors possess the attention that ZnO has received mainly due to its unique chemical and physical properties.1−5 ZnO exhibits a wide bandgap (3.37 eV), high intrinsic electron mobility (∼300 cm2 V−1 s−1), and high exciton binding energy (60 meV).5 ZnO is an n-type semiconductor, which can be converted to a p-type by doping elements such as N and Na.6−8 ZnO is also well renowned for its long-term stability under extreme conditions in addition to its biocompatibility and biosafety.9,10 Due to these unique properties, inexpensive material cost, and high abundance with properly manipulated synthetic strategies and device engineering, ZnO has been successfully used as a cost-effective material in a number of diverse applications including photocatalysis, photovoltaics, sensors, active antibacterial surfaces, optoelectronics, and batteries.11−20 It has always been a challenging task to obtain well-defined 3D architectures of ZnO with a proper understanding of the crystal growth. Over the years, a significant amount of research efforts has been dedicated to developing facile synthetic routes to obtain different 3D nanostructures of ZnO.1−5,21−24 © XXXX American Chemical Society

Compared to one-dimensional structures such as rods and wires and two-dimensional structures such as thin films and disks, 3D nanostructures exhibit unique features such as high surface area, low aggregation, and superior carrier mobility, which can serve well in photocatalysis and sensing applications. Common 3D nanostructures studied in the literature include tetrapods,25,26 hollow27,28 and nonhollow spheres,29 and flower-like structures,14,23,30−32 which are synthesized using different strategies such as hydrothermal, template-assisted aging, and chemical vapor deposition.25−32 General approaches to obtain 3D architectures include the utilization of template directional growth29 or selective manipulation of growth rates of specific crystal facets using surfactants.33−35 For example, Sun et al. developed a hydrothermal route where a zinc amino complex is autoclaved at 180 °C for 12 h to obtain a sea urchin-like 3D architecture for gas sensing applications.36 However, a detailed mechanistic understanding of the growth of these nanostructures can be limited by the growth technique. For instance, it can be Received: June 14, 2019 Accepted: July 26, 2019

A

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 1. TEM (a, b) and SEM (c, d) images of ZnO seeds.

Figure 2. TEM images (a−d) showing the sequential growth of ZnO nanoparticles to form seed clusters at different time intervals of 5, 15, 30, and 60 min, respectively, from adding the base to the zinc acetate solution at 60 °C. (e) UV−vis spectra corresponding to the solutions used for images (a−d).

processes of the SU nanostructures is reported. Furthermore, the photocatalytic activity of the SU ZnO nanostructures is evaluated based on the photodegradation of three organic dyes, indigo carmine (IC), rhodamine B (Rh B), and methylene blue (MB). The results provide evidence for the superior photocatalytic efficiency of the SU ZnO nanostructures.

laborious to track the sequential growth of a nanostructure if the hydrothermal technique is used. Based on the World Bank estimation, 17−20% of contributors for the water pollution is related to the textile industry.37 Annual production of textile dyes is almost a million tons, and around 15% of it is generally released to natural water bodies with minimal or no treatment.37,38 Photodegradation of organic dyes is one of the most widely studied and important areas of photocatalysis, which has the advantages of the usage of costless, inexhaustible, and clean solar energy.39 Semiconductor metal oxides, such as ZnO, an efficient photocatalyst, can be used as a cost-effective tool for pollution control due to the unique chemical and physical properties mentioned earlier. In this work, we present a simple two-step aqueous phase synthetic route to obtain 3D sea urchin (SU)-shaped ZnO nanostructures. Compared to the reported techniques, this synthetic strategy does not involve harsh growth conditions or expensive instruments and only requires 2 h to be completed.40 More importantly, a mechanistic study of the sequential formation and growth

2. RESULTS AND DISCUSSION 2.1. Catalyst Characterization. Transmission electron microscopy (TEM) and scanning electron microscopy (SEM) images of the synthesized ZnO seeds are shown in Figure 1. The TEM images (Figure 1a,b) suggest that the ZnO seeds are clusters of small rod-shaped ZnO nanoparticles. The small ZnO nanoparticles in the size range of 4−6 nm are aggregated to form spherical seed clusters, which are around 80 nm in size, as shown in Figure 1c,d. To gain insight on the mechanism of the formation of the ZnO seeds, TEM images of the growing seeds were recorded at different time intervals (5, 15, 30, 60 min), as shown in Figure 2a−d. After adding the base to zinc acetate, the solution B

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 3. TEM images showing the sequential growth of SU ZnO at different stages, 0 min (a), 15 min (b), 30 min (c), 60 min (d), and UV/vis spectra of the SU ZnO at different growth time intervals, as shown in (e).

(Figure S1). Without CTAB, the absorption edge of the ZnOaggregated clusters increased to 350 nm after 1 h of heating at 60 °C, as shown in Figure 2e. However, in the presence of CTAB, the absorption edge of the ZnO seeds increased only to 336 nm from the original value of 321 nm (Figure S1e) indicating a minimal aggregation of the nanocrystal seeds and a small increase in particles’ size. To compare the growth patterns of the ZnO seeds in the absence and presence of CTAB, both types of seeds were incubated at 90 °C in a solution containing Zn(NO3)2·6H2O and HMTA. After the thermal hydrolysis of HMTA, NH3 is slowly released to the solution which then hydrolyzes in the presence of water to form NH4(OH).42 As the solution becomes basic, Zn2+ ions react with OH− to form ZnO, as shown in eqs 1−3.

became cloudy in about 5 min indicating the formation of the ZnO seeds, as shown in the TEM image in Figure 2a. In the absence of a capping agent in the solution, the ZnO nanoparticles started to aggregate under the constant heating at 60 °C, as shown in the TEM image in Figure 2b recorded after 15 min from adding the base. The aggregation of the individual nanoparticles continues, and a large number of small individual nanoparticles started to be replaced by a smaller number of larger aggregated clusters, as shown in TEM image in Figure 2c recorded after 30 min from adding the base. Finally, the growth process of the seed aggregates appeared to be terminated after 60 min as shown in the TEM image in Figure 2d where the small individual ZnO nanoparticles were totally replaced by a few aggregated clusters in the size range of 80−100 nm. The disappearance of the initially formed small ZnO nanocrystals and the formation of the larger aggregated clusters are a direct consequence of the Ostwald ripening mechanism where the small nanocrystals attain a lower-energy state by transformation into large aggregated clusters with reduced surface to volume ratios.41 The growth of the seed clusters was also studied by recording the UV−vis absorption spectra at different times, as shown in Figure 2e. The absorption edge at 334 nm due to electronic transitions between the valence and conduction bands in the initially formed nanocrystals (5 min) is significantly blue-shifted from the absorption of bulk ZnO, which absorbs at approximately 385 nm (3.2 eV).5 The absorption edge is red-shifted from 334 nm (3.71 eV) to 350 nm (3.54 eV) after the complete growth of the aggregated clusters (60 min) indicating a decrease in bandgap due to increasing particle size. The increase in the bandgap of the small seed clusters is attributed to the quantum confinement mechanism and the seed clusters grow to larger sizes, the bandgap decreases toward the bulk bandgap. These aggregated seed clusters are stable as a suspension in ethanol as confirmed by the TEM images (similar to the image shown in Figure 2d) taken after being suspended in ethanol for 1 week. The main hypothesis in the present synthesis approach is that the formation of ZnO-aggregated seed clusters (Figure 2d) is the key requirement for the secondary growth process to form the unique “sea urchin”-shaped ZnO nanostructures. To test this hypothesis, a capping agent cetyl trimethyl ammonium bromide (CTAB) was added to the zinc acetate solution to prevent the aggregation of the initially formed ZnO nanocrystals into the aggregated seed clusters. As expected, the very small ZnO nanocrystals remained separated and only exhibited a small degree of growth after 1 h of heat treatment at 60 °C

C6H12N4 + 6H 2O → 4NH3 + 6HCHO

(1)

NH3 + 6H 2O F NH+4 + OH−

(2)

Zn 2 + + 2OH− F ZnO + H 2O

(3)

In the case of uncapped ZnO seeds (no CTAB), the growth of the aggregated clusters results in the formation of ZnO rods, which grow along the [0001] direction in a pattern where all of the rods are concentric to the core of the seed cluster yielding the sea urchin-like shape, as shown in the TEM images in Figure 3 recorded at different time intervals of the growth process at 90 °C. Due to the polar nature of the ZnO lattice structure, the surface can be either negatively or positively charged depending on the crystal structure of the terminated plane. Zn-terminated planes will be positively charged, whereas O-terminated planes will be negatively charged. Depending on the charge, the surface will attract counterions in the medium (OH− or Zn2+) and form a layer of alternating charges on the surface resulting in an ordered ZnO lattice.42 This sequence will continue to form an ordered ZnO lattice in the [0001] direction until the concentration of the precursor ions in the solution is depleted. Each nanoparticle on the surface of the seed clusters shown in Figure 3a has the potential to grow into a rod, and this is shown clearly in Figure 3b,c where the rods are grown on the surface of the seed cluster. At a longer growth time, the grown rods become longer and sharper before the growth is terminated by the depletion of the ions in the growth solution as shown in the TEM image in Figure 3d recorded after 60 min during the growth process. The absorption edge of the developed SU ZnO nanostructures exhibits a small red C

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 4. TEM images (a, b) and SEM images (c, d) of the SU ZnO nanostructures.

confirms the suggested mechanism for the growth of the SU structures. To further confirm the hexagonal wurtzite structure of the SU ZnO nanostructures, the Raman spectra of the SU ZnO and the ZnO rods (formed in the presence of CTAB) are shown in Figure 6. In both cases, the prominent peak at 437

shift to 370 nm compared to the original value of 355 nm of the seed clusters, as shown in Figure 3e. Figure 4 displays TEM and SEM images of the full-grown SU ZnO nanostructures. The size distribution of the SU nanostructures ranges from 0.5 to 1.5 μm most likely due to the stagnant growth in the absence of stirring. Interestingly, the growth of the CTAB-capped ZnO seeds did not result in the formation of SU structures but only separated rods with average aspects ratios of ∼10, as shown in Figure S2 (Supporting Information). This result supports the hypothesis that the formation of the aggregated seed clusters shown in Figure 3a is necessary for the growth of the SU structures. It appears that these aggregated seed clusters act as efficient heterogeneous nuclei for the addition of ZnO units to develop the 3D sea urchin structures. Figure 5a,b displays the X-ray diffraction (XRD) patterns of the grown SU nanostructures and the seed clusters of ZnO,

Figure 6. Raman spectra for SU ZnO and ZnO rods.

cm−1 corresponds to the nonpolar E2 (high) optical phonon mode, and the 378 and 580 cm−1 are attributed to the polar transverse optical (TO) A1 and longitudinal optical (LO) E1 phonon modes, respectively. 7 The peak at 332 cm −1 corresponds to the 2E2 mode. According to the selection rules, Raman active phonon resonance modes for the ZnO wurtzite structure are A1 + 2E2 + E2.43 Therefore, the Raman data shown in Figure 6 confirm that the SU ZnO nanostructures have the hexagonal wurtzite crystal structures. 2.2. Photocatalytic Activity. The photocatalytic activity of the SU ZnO nanostructures was evaluated based on the photodegradation efficiency of three organic dyes IC, MB, and Rh B under UV irradiation. The results are shown in Figure 7, which displays the UV−visible spectra of Rh B, IC, and MB recorded at different time intervals during the photodegradation experiments. The degradation percentages of the

Figure 5. XRD patterns of SU ZnO (a), ZnO seeds (b) and TEM images showing how the seeds transformed into SU-shaped ZnO.

respectively. In both cases, the crystal structure belongs to the hexagonal wurtzite structure of bulk ZnO with the space group P63mc. However, the ZnO seed clusters (Figure 5b) show broader XRD peaks compared to the SU ZnO (Figure 5a) due to the smaller particle size. Also, the peak intensity representing the (002) plane is relatively small in the seeds, and it increases significantly in the SU structures, which D

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 7. Photocatalytic activity of SU ZnO in degrading organic dyes (a) Rh B, (b) IC, (c) MB and (d) the apparent rate constants for the photodegradation of Rh B, IC, and MB.

Figure 8. Comparison of photocatalytic activities of SU ZnO and ZnO rods prepared using CTAB-capped ZnO seeds in photodegrading Rh B. (a, b) UV−visible spectra of Rh B showing photodegradation by SU ZnO and ZnO rods, respectively, and (c, d) kinetic plots of the photodegradation reaction by SU ZnO and ZnO rods, respectively.

the 0 min mark of Figure 7c. Weak electrostatic adsorption of MB could shorten its retention time on the surface of the SU ZnO catalyst, which would decrease its photodegradation efficiency. On the other hand, a significant drop in the absorption peak intensity was observed for the Rh B and IC dyes after the initial dark reaction as shown in Figure 7a,b indicating strong electrostatic adsorption on the catalyst surface, and, therefore, a rapid photodegradation was observed within 1 h irradiation in both cases. Finally, the photocatalytic activities of the SU ZnO nanostructures and the ZnO rods in the photodegradation of the Rh B dye were compared. As shown in Figure 8a,b, the photocatalytic activity of ZnO rods is significantly lower than

Rh B, IC, and MB dyes after UV irradiation for 1 h were determined as 96, 99, and 86%, respectively. The photodegradation reactions for all of the three dyes follow pseudofirst-order kinetics (Figure 7d), and the degradation of IC shows the fastest kinetics with a rate constant of 0.076 min−1. According to the data, MB exhibits a slight resistance to photodegradation by the SU ZnO nanostructures compared to the other two dyes. MB is the only anionic dye among the three dyes, and SU ZnO surfaces are expected to possess a slightly negative charge in aqueous solutions due to the surface hydrolysis.44 On account of the electrostatic repulsion, MB showed a little drop in the absorption peak after the initial adsorption−desorption equilibrium for 1 h, which is evident at E

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

the growth solution were mixed together and sonicated for 15 min to disperse the seeds well in the solution. Then, the mixture was incubated in a water bath at 90 °C for 1 h without stirring. Finally, the nanorod clusters were formed and separated from the growth solution and washed with deionized water and ethanol separately to remove the growth solution. The white powder of the sea urchin-shaped ZnO was then oven-dried at 70 °C overnight. The prepared zinc oxide samples were characterized using the following techniques. Transmission electron microscope (TEM) images were obtained using a Jeol JEM-1230 TEM with the Gatan Orius SC1000 side mount CCD camera at 120 kV. Scanning electron microscope (SEM) images were obtained with a Hitachi SU-70 FE-SEM. UV−visible absorbance spectra were obtained using an HP-8453 spectrophotometer. Powder X-ray diffraction (XRD) patterns were acquired with the PANalytical MPD X’Pert Pro with a copper filter (Cu Kα, λ = 1.5405 Å) at 45 kV and 40 mA with a scan speed of 0.5 2θ/min. Raman spectra were obtained from Thermo Scientific DXR SmartRaman (532 nm). A micromeritics 3Flex surface characterization analyzer was used for the surface area measurements. Samples were activated overnight at 100 °C, and the nitrogen adsorption−desorption isotherms were obtained at 77 K to determine the BET surface area. 4.2. Photocatalytic Experiments. Photocatalytic activities of the prepared ZnO samples were evaluated based on the photodegradation of three organic dyes, MB, Rh B, and IC, and the photocatalytic experiment is briefly described as follows. A 4 W, 365 nm UV lamp was used as the source of irradiation, and a Petri dish (diameter of 3.5 in.) was used as the reaction container. The UV lamp was placed above the Petri dish irradiated with the UV light at 365 nm. Before the irradiation experiment, 50 mL of the aqueous solution containing dispersed ZnO nanostructures (0.25 mg/mL) and the dye (10 ppm) was added into the container. Then, the solution was kept in the dark for 1 h to achieve adsorption− desorption equilibrium. Finally, the lamp was turned on, and the photocatalytic reaction was started. Samples were taken at different time intervals to analyze the concentration variation of the dyes. The photodegradation efficiencies were calculated based on the UV−vis absorption of the dyes at different irradiation times. 4.3. Analytical Methods. The concentrations of the dyes were determined by absorbance measurements due to their strong absorption bands in the visible region and the linear correlation between the dye concentration and the absorbance. The absorption peaks of the samples were recorded by a UV− vis spectrophotometer and each measurement was repeated at least three times, and the averaged absorbance values are used to calculate the dye concentrations.

that of the SU nanostructures. As shown in Figure 8c,d, after 1 h of UV irradiation, ZnO rods only degraded 34% of the dye (10 ppm), whereas the SU nanostructures degraded 96% of the dye. The increased activity of the SU nanostructures could be attributed to the BET specific surface area of 26.9 m2/g, which is significantly larger than that of ZnO rods (17.7 m2/g).45,46 The relevant N2 adsorption/desorption isotherms of the SU ZnO nanostructures and the ZnO rods are shown in Figure S3 (Supporting Information). The enhanced photocatalytic activity could also be attributed to the porous structure of the SU ZnO nanostructures, which would enhance the adsorption and probably the trapping of the organic dyes within the assembly of the short rods for efficient photodegradation.

3. CONCLUSIONS In summary, a simple approach has been developed for the synthesis of sea urchin nanostructures by the formation and assembly of ZnO nanorods under favorable growth conditions in an aqueous solution. The thermal treatment of a basic zinc acetate solution in ethanol results in the formation of aggregated seed clusters consisting of small ZnO nanorods, which were then grown in a precursor solution containing Zn(NO3)2·6H2O and hexamethylenetetramine to assemble the SU structures from the anisotropic ZnO nanorods on the surface of the seed clusters. Each ZnO nanoparticle in the aggregated seed clusters grew sequentially into a ZnO nanorod, and the nanorods were concentric to the core of the clusters yielding the unique SU-like shape. In the presence of a capping agent such as CTAB, the aggregated seed clusters were not formed, and the growth of the CTAB-capped ZnO nanorods resulted in separated rods with average aspect ratios of ∼10. The SU ZnO nanostructures exhibit a hexagonal wurtzite crystal structure and higher specific surface area (26.9 m2/g) than the CTAB-capped nanorods (17.7 m2/g). The SU ZnO nanostructures show superior photocatalytic efficiency for the degradation of three common organic dyes compared to the ZnO nanorods. The removal efficiencies of indigo carmine, methylene blue, and rhodamine B by the SU nanostructures were 99, 86, and 96%, respectively, after 1 h of UV irradiation. Therefore, the ZnO SU structures have the potential to be a versatile photocatalyst for the photodegradation of organic dyes in industrial wastewater. 4. EXPERIMENTAL SECTION 4.1. Preparation and Characterization of Sea UrchinShaped ZnO Nanorod Clusters. Zinc oxide seeds were prepared in a colloidal solution using ethanol as the solvent according to the method described in ref 39. In a typical synthesis, 2 mM zinc acetate [Zn(Ac)2, Alfa Aesar] in ethanol was heat-treated for 30 min at 70 °C under stirring in a water bath. Then, 4 mM potassium hydroxide [KOH, Sigma] in ethanol was added dropwise. The molar ratio between Zn(Ac)2 and KOH was maintained at 1:1. Then, the temperature was reduced to 60 °C with continued heating for 1 h under constant stirring. The reaction solution started to become cloudy with time indicating the formation of the ZnO seeds. To synthesize sea urchin-shaped ZnO nanorod clusters, an equimolar aqueous solution of zinc nitrate hexahydrate [Zn(NO3)2·6H2O, Sigma-Aldrich] and hexamethylenetetramine [HMTA, Sigma] (20 mM) was used as the growth solution. First, 25 mL of the ZnO seed solution and 100 mL of



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01772. TEM images showing the sequential growth of ZnO seeds in the presence of CTAB at different time intervals (Figure S1); TEM images of ZnO rods synthesized with CTAB-capped ZnO seeds and the UV−vis spectrum (Figure S2); and N2 adsorption/desorption isotherms at F

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega



Article

Preparation for Dye Sensitized Solar Cells. Langmuir 2011, 27, 3920−3929. (14) Ahmad, M.; Yingying, S.; Nisar, A.; Sun, H.; Shen, W.; Wei, M.; Zhu, J. Synthesis of Hierarchical Flower-like ZnO Nanostructures and their Functionalization by Au Nanoparticles for Improved Photocatalytic and High Performance Li-ion Battery Anodes. J. Mater. Chem. 2011, 21, 7723−7729. (15) Alenezi, M. R.; Alshammari, A. S.; Alzanki, T. H.; Jarowski, P.; Henley, S. J.; Silva, S. R. P. ZnO Nanodisk Based UV Detectors with Printed Electrodes. Langmuir 2014, 30, 3913−3921. (16) Wang, X.; Chen, K.; Zhang, Y.; Wan, J.; Warren, O. L.; Oh, J.; Li, J.; Ma, E.; Shan, Z. Growth Conditions Control the Elastic and Electrical Properties of ZnO Nanowires. Nano Lett. 2015, 15, 7886− 7892. (17) Jeem, M.; Zhang, L.; Ishioka, J.; Shibayama, T.; Iwasaki, T.; Kato, T.; Watanabe, S. Tuning Optoelectrical Properties of ZnO Nanorods with Excitonic Defects via Submerged Illumination. Nano Lett. 2017, 17, 2088−2093. (18) Pelicano, C. M.; H. Yanagi, H. Efficient Solid-State Perovskite Solar Cells Based on Nanostructured Zinc Oxide Designed by Strategic Low Temperature Water Oxidation. J. Mater. Chem. C 2017, 5, 8059−8070. (19) Pelicano, C. M.; Yanagi, H. pH-Controlled Surface Engineering of Nanostructured ZnO Films Generated via a Sustainable Lowtemperature H2O Oxidation Process. Appl. Surf. Sci. 2019, 467−468, 932−939. (20) Pelicano, C. M.; Yanagi, H. Enhanced Charge Transport in Aldoped ZnO Nanotubes Designed via Simultaneous Etching and Al Doping of H2O-oxidized ZnO Nanorods for Solar Cell Applications. J. Mater. Chem. C 2019, 7, 4653−4661. (21) Cho, S.; Jang, J.-W.; Lee, J. S.; Lee, K.-H. Exposed Crystal Face Controlled Synthesis of 3D ZnO Superstructures. Langmuir 2010, 26, 14255−14262. (22) Song, B.; Wang, Y.; Cui, X.; Kou, Z.; Si, L.; Tian, W.; Yi, C.; Wei, T.; Sun, Y. A Series of Unique Architecture Building of Layered Zinc Hydroxides: Self-Assembling Stepwise Growth of Layered Zinc Hydroxide Carbonate and Conversion into Three-Dimensional ZnO. Cryst. Growth Des. 2016, 16, 887−894. (23) Umar, A.; Lee, S.; Im, Y. H.; Hahn, Y. B. Flower-shaped ZnO Nanostructures obtained by Cyclic Feeding Chemical Vapour Deposition: Structural and Optical Properties. Nanotechnology 2005, 16, 2462. (24) Resmini, A.; Tredici, I. G.; Cantalini, C.; Giancaterini, L.; De Angelis, F.; Rondanina, E.; Patrini, M.; D. Bajoni, D.; AnselmiTamburini, U. A Simple all-Solution Approach to the Synthesis of lLarge ZnO Nanorod Networks. J. Mater. Chem. A 2015, 3, 4568− 4577. (25) Lupan, O.; Postica, V.; Gröttrup, J.; Mishra, A. K.; de Leeuw, N. H.; Carreira, J. F. C.; Rodrigues, J.; Ben Sedrine, N.; Correia, M. R.; Monteiro, T.; Cretu, V.; Tiginyanu, I.; Smazna, D.; Mishra, Y. K.; Adelung, R. Hybridization of Zinc Oxide Tetrapods for Selective Gas Sensing Applications. ACS Appl. Mater. Interfaces 2017, 9, 4084− 4099. (26) Gedamu, D.; Paulowicz, I.; Kaps, S.; Lupan, O.; Wille, S.; Haidarschin, G.; Mishra, Y. K.; Adelung, R. Rapid Fabrication Technique for Interpenetrated ZnO Nanotetrapod Networks for Fast UV Sensors. Adv. Mater. 2014, 26, 1541−1550. (27) Yu, J.; Yu, X. Hydrothermal Synthesis and Photocatalytic Activity of Zinc Oxide Hollow Spheres. Environ. Sci. Technol. 2008, 42, 4902−4907. (28) Wang, X.; Hu, P.; Fangli, Y.; Yu, L. Preparation and Characterization of ZnO Hollow Spheres and ZnO−Carbon Composite Materials Using Colloidal Carbon Spheres as Templates. J. Phys. Chem. C 2007, 111, 6706−6712. (29) Yao, K. X.; Zeng, H. C. ZnO/PVP Nanocomposite Spheres with Two Hemispheres. J. Phys. Chem. C 2007, 111, 13301−13308. (30) Li, J.; Lu, G.; Wang, Y.; Guo, Y.; Guo, Y. A High Activity Photocatalyst of Hierarchical 3D Flowerlike ZnO Microspheres:

77 K for the ZnO sea urchin and nanorod structures (Figure S3) (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

M. Samy El-Shall: 0000-0002-1013-4948 Present Address

† School of Environmental and Municipal Engineering, Xi’an University of Architecture and Technology, Xi’an, Shaanxi 710055, P. R. China (L.B.).

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Science Foundation through grant CHE-1463989. The authors would like to thank Karthik V. Pulavarthi from Maggie L. Walker Governor’s School, Richmond, Virginia for his involvement in this project.



REFERENCES

(1) Li, J.; Fan, H.; Jia, X. Multilayered ZnO Nanosheets with 3D Porous Architectures: Synthesis and Gas Sensing Application. J. Phys. Chem. C 2010, 114, 14684−14691. (2) Cho, S.; Jang, J.-W.; Kim, J.; Lee, J. S.; Choi, W.; Lee, K.-H. Three-Dimensional Type II ZnO/ZnSe Heterostructures and Their Visible Light Photocatalytic Activities. Langmuir 2011, 27, 10243− 10250. (3) Liu, J.; Huang, H.; Zhao, H.; Yan, X.; Wu, S.; Li, Y.; Wu, M.; Chen, L.; Yang, X.; Bao-Lian Su, B.-L. Enhanced Gas Sensitivity and Selectivity on Aperture-Controllable 3D Interconnected Macro− Mesoporous ZnO Nanostructures. ACS Appl. Mater. Interfaces 2016, 8, 8583−8590. (4) Katea, S. N.; Hajduk, Š .; Orel, Z. C.; Westin, G. Low Cost, Fast Solution Synthesis of 3D Framework ZnO Nanosponges. Inorg. Chem. 2017, 56, 15150−15158. (5) Wang, Z. L. Zinc Oxide Nanostructures: Growth, Properties and Applications. J. Phys.: Condens. Matter 2004, 16, R829−R858. (6) Lu, J. G.; Ye, Z. Z.; Zhuge, F.; Zeng, Y. J.; Zhao, B. H.; Zhu, L. P. p-Type Conduction in N−Al co-Doped ZnO Thin Films. Appl. Phys. Lett. 2004, 85, 3134−3135. (7) Herring, N. P.; Panchakarla, L. S.; El-Shall, M. S. P-type Nitrogen-Doped ZnO Nanostructures with Controlled shape and Doping Level by Facile Microwave Synthesis. Langmuir 2014, 30, 2230−2240. (8) Liu, W.; Xiu, F.; Sun, K.; Xie, Y.-H.; Wang, K. L.; Wang, Y.; Zou, J.; Yang, Z.; Liu, J. Na-Doped p-Type ZnO Microwires. J. Am. Chem. Soc. 2010, 132, 2498−2499. (9) Li, Z.; Yang, R.; Yu, M.; Bai, F.; Li, C.; Wang, Z. L. Cellular Level Biocompatibility and Biosafety of ZnO Nanowires. J. Phys. Chem. C 2008, 112, 20114−20117. (10) Hong, H.; Wang, F.; Zhang, Y.; Graves, S. A.; Eddine, S. B. Z.; Yang, Y.; Theuer, C. P.; Nickles, R. J.; Wang, X.; Cai, W. Red Fluorescent Zinc Oxide Nanoparticle: A Novel Platform for Cancer Targeting. ACS Appl. Mater. Interfaces 2015, 7, 3373−3381. (11) Su, Y. K.; Peng, S. M.; Ji, L. W.; Wu, C. Z.; Cheng, W. B.; Liu, C. H. Ultraviolet ZnO Nanorod Photosensors. Langmuir 2010, 26, 603−606. (12) Brayner, R.; Dahoumane, S. A.; Yepremian, C.; Djediat, C.; Meyer, M.; Coute, A.; Fievet, F. ZnO Nanoparticles: Synthesis, Characterization, and Ecotoxicological Studies. Langmuir 2010, 26, 6522−6528. (13) Marczak, R.; Werner, F.; Ahmad, R.; Lobaz, V.; Guldi, D. M.; Peukert, W. Detailed Investigations of ZnO Photoelectrodes G

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Synthesis, Characterization and Catalytic Activity. J. Colloid Interface Sci. 2012, 377, 191−196. (31) Rahman, Q. I.; Ahmad, M.; Misra, S. K.; Lohani, M. B. Hexagonal ZnO Nanorods Assembled Flowers for Photocatalytic Dye Degradation: Growth, Structural and Optical Properties. Superlattices Microstruct. 2013, 64, 495−506. (32) Li, B.; Wang, Y. Facile Synthesis and Enhanced Photocatalytic Performance of Flower-like ZnO Hierarchical Microstructures. J. Phys. Chem. C 2010, 114, 890−896. (33) Bahng, J. H.; Yeom, B.; Wang, Y.; Tung, S. O.; Hoff, J. D.; Kotov, N. Anomalous Dispersions of “hedgehog” Particles. Nature 2015, 517, 596−599. (34) Ge, J.; Bo, T.; Linhai, Z.; Zhiqiang, S. A Rapid Hydrothermal Route to Sisal-like 3D ZnO Nanostructures via the Assembly of CTA+ and Zn(OH)42−: Growth Mechanism and Photoluminescence Properties. Nanotechnology 2006, 17, 1316. (35) Herring, N. P.; AbouZeid, K.; Mohamed, M. B.; Pinsk, J.; ElShall, M. S. Formation Mechanisms of Gold−Zinc Oxide Hexagonal Nanopyramids by Heterogeneous Nucleation using Microwave Synthesis. Langmuir 2011, 27, 15146−15154. (36) Sun, Y.; Hu, J.; Wang, N.; Zou, R.; Wu, J.; Song, Y.; Chen, H.; Chen, H.; Chen, Z. Controllable Hydrothermal Synthesis, Growth Mechanism, and Properties of ZnO Three-Dimensional Structures. New J. Chem. 2010, 34, 732−737. (37) Kant, R. Textile Dyeing Industry an Environmental Hazard. Nat. Sci. 2012, 04, 22−26. (38) Ajmal, A.; Majeed, I.; Malik, R. N.; Idriss, H.; Nadeem, M. A. Principles and Mechanisms of Photocatalytic Dye Degradation on TiO2-based Photocatalysts: A Comparative Overview. RSC Adv. 2014, 4, 37003−37026. (39) Baruah, S.; Jaisai, M.; Imani, R.; Nazhad, M. M.; Dutta, J. Photocatalytic Paper using Zinc Oxide Nanorods. Sci. Technol. Adv. Mater. 2010, 11, No. 055002. (40) Karlsson, M. E.; Mamie, Y. C.; Calamida, A.; Gardner, J. M.; Ström, V.; Pourrahimi, A. M.; Olsson, R. T. Synthesis of Zinc Oxide Nanorods via the Formation of Sea Urchin Structures and Their Photoluminescence after Heat Treatment. Langmuir 2018, 34, 5079− 5087. (41) Ratke, L.; Voorhees, P. W. Growth and Coarsening: Ostwald Ripening in Material Processing; Springer Science & Business Media, 2002; ISBN: 3540425632, 9783540425632. (42) Li, Q.; Kumar, V.; Li, Y.; Zhang, H.; Marks, T. J.; Chang, R. P. H. Fabrication of ZnO Nanorods and Nanotubes in Aqueous Solutions. Chem. Mater. 2005, 17, 1001−1006. (43) Jiang, P.; Zhou, J. J.; Fang, H. F.; Wang, C. Y.; Wang, Z. L.; Xie, S. S. Hierarchical Shelled ZnO Structures Made of Bunched Nanowire Arrays. Adv. Funct. Mater. 2007, 17, 1303−1310. (44) Degen, A.; Kosec, M. Effect of pH and Impurities on the Surface Charge of Zinc Oxide in Aqueous Solution. J. Eur. Ceram. Soc. 2000, 20, 667−673. (45) Zhang, X.; Qin, J.; Hao, R.; Wang, L.; Shen, X.; Yu, R.; Limpanart, S.; Ma, M.; Liu, R. Carbon-Doped ZnO Nanostructures: Facile Synthesis and Visible Light Photocatalytic Applications. J. Phys. Chem. C 2015, 119, 20544−20554. (46) Hariharan, C. Photocatalytic Degradation of Organic Contaminants in Water by ZnO Nanoparticles: Revisited. Appl. Catal., A 2006, 304, 55−61.

H

DOI: 10.1021/acsomega.9b01772 ACS Omega XXXX, XXX, XXX−XXX