You are on page 1of 37

Accepted Manuscript

Title: ZnO nanotubes: Preparation and photocatalytic


performance evaluation

Authors: Parvin Samadipakchin, Hamid Reza Mortaheb,


Alireza Zolfaghari

PII: S1010-6030(15)30196-9
DOI: http://dx.doi.org/doi:10.1016/j.jphotochem.2017.01.018
Reference: JPC 10506

To appear in: Journal of Photochemistry and Photobiology A: Chemistry

Received date: 18-10-2015


Revised date: 26-11-2016
Accepted date: 11-1-2017

Please cite this article as: Parvin Samadipakchin, Hamid Reza Mortaheb,
Alireza Zolfaghari, ZnO nanotubes: Preparation and photocatalytic performance
evaluation, Journal of Photochemistry and Photobiology A: Chemistry
http://dx.doi.org/10.1016/j.jphotochem.2017.01.018

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
 ZnO nanotube is synthesized by a simple reproducible solution-based
method.

 Reaction time and temperature are found to play key roles in synthesis of
nanotubes.

 ZnO nanotubes represent higher photocatalytic stability than ZnO


nanoparticles.
ZnO nanotubes: preparation and photocatalytic performance evaluation

Parvin Samadipakchin, Hamid Reza Mortaheb*, Alireza Zolfaghari

Chemistry and Chemical Engineering Research Center of Iran, Tehran, Iran

ABSTRACT

In this work, uniform zinc oxide nanotube was successfully synthesized by a simple
reproducible solution-based method. The effects of temperature, molecular weight of
polyethylene glycol, and synthesis time on morphology and photocatalytic properties of the
products were investigated. The synthesized particles were characterized by X-ray
diffraction, electron microscopy, laser particle sizing, thermal gravimetric analysis,
differential scanning calorimetry, and photoluminescence spectrum. The reaction temperature
and reaction time were found to play key roles in synthesis of the nanotubes. The effects of
annealing on crystallinity, optical property, and photocatalytic activity of the nanotubes were
also investigated. The photocatalytic activity of nanotubes was evaluated by degrading Acid
Blue 9 as a pollutant dye in a slurry reactor under UV light. The photocatalytic activity of the
synthesized ZnO nanotube remains almost unchanged after three times reusing that indicates
its improved stability compared to that of a commercial ZnO nanoparticle.

Keywords
ZnO nanotube; Photocatalyst; Polluting dye; Annealing

*
Corresponding author

* Tel: +98-21-44787751, Fax: +98-21-44787781, E-mail: mortaheb@ccerci.ac.ir

1
1. Introduction

Semiconductors have drawn much attention in recent years owing to their interesting

applications in electronics, solar energy conversion, environmental purification, and

production of sensors and transducers [1-3]. These materials are capable to degrade organic

contaminants in water by photodegradation processes [4]. Among various

semiconductors,TiO2 and ZnO have been studied more extensively because of their strong

oxidizing power and non-toxicity nature [5] where zinc oxide is the more promising

semiconductor with growing applications because of its advantages such as low cost and that

it adsorbs UV light over a larger spectrum than TiO2 [6]. Zinc oxide is a II–VI semiconductor

with a wide and direct band gap of about 3.37 eV (at 300 K), a large free exciton binding

energy of 60meV[7], and high mechanical and thermal stabilities [8].

The chemical and physical properties of nanosized materials are different from those of the

bulk materials [9]. Their high specific surface area disposes their structural defects into the

surface. These characteristics make nanocatalysts more appropriate for industrial applications

[6]. Tubular forms of nanosized materials have particularly attractive because of their lower

densities and 1-dimensional geometry, which facilitate fast electron transporting for longer

distances. 1-dimensional nanostructures have a larger specific surface area and pore volume.

Also, because of their high length-to-diameter ratio of these nanostructures, the light

absorption and scattering are significantly enhanced [10-12].

ZnO nanotubes are usually synthesized by thermal evaporation [13], microwave plasma

deposition [14], electrodeposition [15], molecular beam epitaxy [16], template-based [17],

hydrothermal [18], and solution-based methods [19]. Most synthesis techniques are difficult

and require high temperatures (> 350◦C) [20]. However, the solution-based methods require

2
low temperature (25-200◦C), relatively simple equipment, mild reaction conditions, and

potential for scaling up [21]. The reaction temperature, nature of reactants and their

concentrations [22], type of surfactant [20], pH [23], and many other parameters can

influence synthesis of ZnO [24-26]. Controlling of all these conditions is a difficult task.

Accordingly, synthesis of ZnO nanotubes by the method has poor repeatability as confirmed

by our primary experimental results. Furthermore, most ZnO synthesis methods, have been

performed in temperatures higher than 100◦C and in an autoclave [27-29]. This requires

performing the process at high pressures in a closed reactor.

Although the effect of different synthesis conditions on the properties of ZnO nanorods have

been studied by several researches, a similar study on synthesis of ZnO nanotubes has not

been done thoroughly [20, 25, 30].

In the present research, ZnO nanotubes are synthesized by a simple and reproducible

solution-based method. The proposed technique can be performed without using any

autoclave or catalyst. The initial reactants are inexpensive and nontoxic. The photocatalytic

performance of the synthesized ZnO nanotubes and the effects of different parameters such as

temperature, molecular weight of the template, and synthesis time are then investigated by

photodegradation of Acid Blue 9 (AB9) as a polluting dye in a slurry reactor.

2. Experimental

2.1. Materials

Zn(NO3)2.6H2O with 99% purity was purchased from Fluka. Ammonia solution (25%) and

PEGs with molecular weights of 1000, 2000, 4000, and 10000 were purchased from Merck.

A Commercial ZnO nanoparticle (mesh size of less than 50 nm) and purity of 99.99% was

3
provided by Neutrino. All the chemicals in this work were of analytical grades and used

without further purifications.

2.2. Synthesis procedure

3 g Zn(NO3)2.6H2O and 0.9 g PEG with the specified molecular weight in each experiment

were added into 375 ml deionized water. The solution was stirred at ambient temperature for

15 min. The solution was placed in an ultrasonic bath and a 1M aqueous solution of NH3was

added drop by drop (rate= 1.1ml/min) to set the pH of solution to 10. The obtained slurry was

finally heated at constant temperature while it was stirred on a heater magnet at 700 rpm until

the reaction was completed. The precipitates were then filtered and washed several times by

distilled water and ethanol to remove nitrate ions and PEG residuals. They were finally dried

in an oven at 60◦C under air atmosphere.

To investigate the effects of reaction temperature, reaction time, and molecular weight of

PEG, the experiments were carried out at 70, 75, 78 and 80◦C for 8, 12, 16 and 24 h with

PEG1000, 2000, 4000 and 10000, respectively.

2.3. Physicochemical characterizations

The crystalline structure and phase purity of ZnO nanotubes were characterized using X-Ray

Diffraction (XRD), Bruker D8 Advance. The morphology of the product was studied using

Scanning Electron Microscopy (SEM), Tescan, Vegall, and Transmission Electron

Microscopy (TEM), Ziess, EM900.The optical properties of the samples were analyzed by

4
Photo Luminescence Spectrophotometer (PLS), Jasco, ST-6500. The spectra were obtained

with an excitation wavelength of 325nm from a xenon emitter at 25◦C.

The thermal analyses of ZnO nanotubes were carried out by Thermo Gravimetric Analysis

(TGA) (NETZSC, TG 209F1) and Differential Scanning Calorimetry (DSC) (NETZSCH,

DSC 204 F1) under air atmosphere with a rate of 10◦C/min.

The specific surface area for the synthesized ZnO nanotubes was determined by analysis of

BET (BELSORP-mini II) data.

The pH of Point of Zero Charge (PZC) of synthesized ZnO nanotubes was obtained by

simultaneous pH-metric and conductometric titrations of the nanotubes in slurry.

The particle size and size distribution of the dispersed ZnO nanotubes were obtained by Laser

Particle Sizing (LPS) apparatus (FRITSCH, Analysette Nanotec). The particle size was

determined by laser irradiation to the suspended nanotubes in water. As the particles could be

aggregated in pHs close to pHPZC and this might cause an error in determination of particles’

sizes, the slurries with different pHs in the range of 5-12 were prepared and their size

distributions were determined by LPS apparatus.

2.4. Photocatalytic tests

In order to evaluate the photocatalytic activity of the synthesized ZnO, the nanostructure was

subjected to the photodegradation of AB9 as a polluting commercial dye. The experiments

were carried out in a slurry-type photoreactor. The details of the reactor and experimental

procedure were explained elsewhere [31]. All experiments were carried out in pH of 7.0 at

constant temperature (25◦C). The suspended feed solution contained 2000 ppm of ZnO and

20 ppm of AB9. The photocatalyst was completely dispersed in the feed by an ultrasonic

bath. A UV lamp with radiation wave length of 254nm was used as the light source. The
5
equilibrium condition in adsorption of AB9 on the surface of catalyst was attained by primary

stirring of the feed in dark conditions for 30 min. The concentrations of AB9 in the samples

taken from the reactor were determined by UV-Vis spectroscopy (at λmax = 630 nm).

3. Results and Discussion

3.1. Effect of process parameters on synthesis and photocatalytic performance of ZnO

Synthesis of ZnO was performed under different conditions as summarized in Table 1. As

seen in the table, ZnO cannot be produced in the temperatures below 78◦C. Also, the reaction

time of less than 16 h is not sufficient for production of ZnO in the form of nanotube. The

ZnO could be synthesized with different molecular weights of PEG. The effects of reaction

temperature, molecular weight of PEG, reaction time on synthesis and photocatalytic

performance of ZnO are discussed in the following sections.

3.1.1. Effect of temperature

The reaction temperature plays an important role in the synthesis and morphology of ZnO

nanostructures.

In general, the synthesis mechanism of zinc oxide in the presence of ammonia is often

proposed based on the following step reactions[18, 32]:

Zn(NO3)2 → Zn2+ + 2NO3- (1)

NH3 + H2O ↔ NH4+ + OH- (2)

Zn2+ + 4 NH3.H2O ↔ [Zn(NH3)4]2+ + 4H2O (3)

[Zn(NH3)4]2+ + 2OH- ↔ ZnO + 4NH3 + H2O (4)

6
Zn2+ + 4OH- ↔ [Zn(OH)4]2- (5)

[Zn(OH)4]2- → ZnO + H2O + 2OH- (6)

In a strong basic condition (pH = 10), in which zinc mainly presents in the form of

[Zn(NH3)4]2+, Zn2+ provided by decomposition of [Zn(NH3)4]2+ produces the growth unit of

[Zn(OH)4]2-. The growth unit is then decomposed to produce ZnO.

However, at high temperature, such as those in the present research, the complexation

reactions are not favored due to entropy effect. In these conditions the majority of zinc is in

the form Zn2+(aq) [33]. Then, zinc oxide can be precipitated from aqueous solution either

directly:

Zn2+ + 2OH-↔ ZnO + H2O (7)

or via a successive reversible reactions by formation of zinc hydroxide as the intermediate:

Zn2+ + 2OH-↔ Zn(OH)2↔ ZnO + H2O (8)

In the latter mechanism, the less-stable phase of Zn(OH)2 is precipitated first and then it

undergoes a reversible reaction to form ZnO. The conversion of Zn(OH)2 to ZnO in the last

step of the successive reactions depends on the temperature and reaction time [34].

Fig. 1a shows the XRD patterns of ZnO nanostructures, which were synthesized at different

reaction temperatures. The XRD patterns clearly show that in the syntheses at 70 and 75◦C,

the peaks of both zinc hydroxide (wulfingite) and zinc oxide (wurtzite) are present while the

ratio peak intensities for ZnO/Zn(OH)2 i.e. the concentration ratio of ZnO/Zn(OH)2 is

increased when the temperature is increased from 70 to 75◦C. Zn(OH)2 is the primary

precipitated as a gel-type precipitant. It is then converted to the crystalline structure of zinc

hydroxide (wulfingite) by heating. The crystalline zinc hydroxide is finally converted to ZnO

due to heating and aging. The observation of crystalline zinc hydroxide in the XRD patterns

may show that the reaction is not completed at 70 and 75◦C within 24 hours. The XRD for the

7
experiment at 75◦C and reaction time of 48 h showed a hexagonal wurtzite ZnO structure

with no impurities.

On the other hand, for the syntheses at 78 and 80◦C, only the wurtzite ZnO with good

crystallinity was detected. All the reflections could be indexed to hexagonal wurtzite

structure of ZnO with lattice parameters a = 3.249Å and c = 5.206 Å, which were in good

agreement with the reported data for wurtzite ZnO (a = 3.249 Å, c = 5.205 Å, JCPDS File 5-

664). The XRD patterns indicate that well-crystallized hexagonal ZnO are obtained under

present conditions. The sharp and narrow diffraction peaks exhibit good crystallinity. No

characteristic peaks from the intermediates such as Zn(OH)2 are detected. The much weaker

intensity of the (002) peak as compared to that in the standard JCPDS card (36-1451)

provides further evidence of the tubular structure as stated in literature [35].

The effects of reaction temperature on morphology of the products are shown in SEM

images of Fig. 2. Each figure contains a SEM image with a larger magnification as a subset.

Fig. 2a corresponds to the synthesis at 75◦C. The image clearly reveals that the product is

composed of particles with diameters up to about 4 μm, and nanorods with a wide range of

diameters (130nm-1μm) and lengths (1 μm-5μm). Fig. 2b shows the morphology of products

synthesized at 78◦C. Uniform tubular structures with average diameter and length of 340 nm

and 3 μm, respectively, can be seen in the figure. The image clearly shows a variable

diameter along the length i.e. a horn-shape with one closed end. The subset image in Fig. 2b

clearly shows the open end of the nanotubes and its hexagonal prismatic structure. Fig. 2c

corresponds to the products synthesized at 80◦C. The figure again shows a tubular structure.

However, compared to the product synthesized at 78◦C, their shapes are more irregular with a

wide range of size distribution for diameter (about 200-370 nm) and length (2-4.5μm).

The comparison between reductions in concentration of AB9 by photocatalytic

degradation using the synthesized nanotubes at 78 and 80◦C as an index of their


8
photocatalytic activities (Fig. 3a) confirms higher activity of the nanotubes synthesized at T=

78◦C. According to the previous researches, the degradation rate of photocatalytic reaction for

most organic compounds at low concentrations can be described by pseudo-first order

kinetics [36-38]. The rate constant can then be obtained by plotting ln(C/C0) versus time. Fig.

3b shows that the rate constants for the nanotubes produced at T= 78 and 80◦C are 0.043 and

0.023 (min-1), respectively. This indicates the photocatalytic activity of about twice for the

nanotubes synthesized at T= 78◦C. The higher activity of nanotubes synthesized at T= 78◦C

may be attributed to two parameters:

1) As the SEM images confirms, the size of nanotubes in T= 78◦C are smaller than the

average size of nanotubes in T= 80◦C. This will increase the specific surface area and

therefore their activity as a photocatalyst.

2) By increasing the synthesis temperature, the defects in the crystalline structure including

surface defects are reduced. In a photocatalytic process, the surface defects can trap

photo-induced electrons to inhibit recombination of photo-induced electrons and holes

[39]. Therefore, high concentration of surface defects may result in higher photocatalytic

activity.

3.1.2. Effect of reaction time

As stated above, the orthorhombic wulfingite Zn(OH)2 is formed as an intermediate product

in the synthesis of ZnO in the present method. Further dehydration of Zn(OH)2 into ZnO can

be performed in an aging process by heating over the time. The process can be accelerated

either by raising temperatures [34] or by increasing synthesis time [40] while the required

reaction time is reduced by increasing reaction temperature.

9
As the properties of any growing nanostructure in a reactive media are time-dependent, in

order to investigate the effect of reaction time, the syntheses were performed in 8, 12, 16, and

24 h at T= 78◦C with PEG 2000 as the soft template. The crystalline structures of the products

were examined by XRD. The results show that all the samples have hexagonal wurtzite

structure with no characteristic peaks of Zn(OH)2 or other impurities.

The SEM images of the synthesized products in 8, 12, and 16h, are shown in Figs. 2d-f,

respectively. Fig. 2d shows that the crystals synthesized in 8h have a nanorod shape with

average diameter of 180 nm and length of 1.6 μm while the crystals synthesized in 12 h, Fig.

2e, are a mixture of nanorods and nanotubes with the mean diameter of 380 nm and length of

4.3 μm. Also, it can be seen in Fig. 2f that the majority of crystals synthesized in 16 h have a

tubular shape with a mean size of 460nm in diameter and 5.2 μm in length. The increasing in

the length and diameter of the crystals by increasing the reaction time as well as alteration in

their morphology form can be interpreted considering the procedure of crystal’s growth:

The growth of ZnO crystals at earlier reaction times may proceed through a range of

dissolution/recrystallization processes. In the wurtzite ZnO structure, the growth rate is faster

in two polar facets (Zn-terminated and O-terminated facets) along the [0001] direction [41].

At earlier reaction times, when the reactant concentrations are high enough, the crystal

growth is progressed uniformly along the [0001] direction. However, as the concentrations of

reactants are decreased during the experiment, supplying the Zn2+/OH- for crystal growth will

be restricted. Therefore, in the Zn-terminated facet, where the positive charge density at the

center of the surface is higher than that near the edges, the concentration of Zn2+ in the double

layer adjacent to the center is lower than that at the edges. Therefore, the growth of crystal is

preferably occurred at the edges of the facet i.e. the nanotube structure is formed. A similar

trend takes place on the O-terminated facet where the negative charge density is higher at the

10
center of the facet than that at the edges, and again crystal growth would be in the form of

nanotube. The reaction may then evolve from nanorod to nanotube structure through the

dissolution/recrystallization mechanism[33].

On the other hand, comparing the lengths and diameters of nanotubes synthesized in

16 h (Fig. 2f) and those synthesized in 24 h (Fig. 2b) indicates that both diameter and length

of nanotubes are reduced by increasing reaction time from 16 h to 24 h. The decrease in

diameter may be due to dissolution of ZnO crystals at the end hours of the reaction period

when the dissolution in the low-concentration solution is dominant comparing to the

recrystallization process. The decrease in the length may be due to either dissolution or

longitudinal rupture.

The rate constants of photocatalytic degradation by nanostructures synthesized in different

reaction times are k24h= 0.043, k16h= 0.053, k12h= 0.050, and k8h= 0.034. Although the surface

areas of the 24-h nanotubes are higher than that of the 16-h ones, the rate constant is

decreased when the synthesis time is increased from 16 h to 24 h. This might be due to the

presence of more structural defects in the 16-h nanotubes as higher synthesis time may allow

better orientation in crystalline structure and reducing the structural defects. However, as the

synthesis time is decreased from 16 h to 12 h, a slight decrease in the rate constant is

observed. This might be due to the outcome of two alternative parameters; the presence of

defects and the change in morphology. The relatively low reaction time of 12 h allows

presence of more structural defects in the product. On the other hand, because both nanorods

and nanotubes are present in the 12-h product, and due to higher density of the nanorods, the

available surface area for the photocatalytic reaction per unit weight of the photocatalyst is

lower i.e. the lower photocatalytic activity of the 12-h product. The result also shows that the

rate constant decreases significantly when the reaction time is decreased from 12 h to 8 h.

11
This is due to lower surface area per unit weight of the photocatalyst in the 8-h product where

nanorod is the only detected structure.

3.1.3. Effect of molecular weight of PEG

The unidirectional synthesis of nanostructures can be assisted by applying some surfactants

as soft templates [42, 43]. The surfactant molecules above their critical micelle

concentrations (CMC) associate with themselves dynamically and spontaneously in

aggregates with special forms such as micelles, lamellas, rod-like micelles, and liposomes

[44]. The mechanism of acting surfactant as a template in synthesis of nanostructures is not

known clearly. PEG as a nonionic liner water-soluble surfactant can influence the

morphology and the anisotropic growth of nanocrystals [14, 18]. It was found that one-

dimensional ZnO crystals cannot be synthesized in the absence of PEG [43].The O atoms of

PEG molecules have strong electrostatic interactions with the Zn2+ ions. Therefore, the long

curly chains of PEG molecules in the solution will embrace Zn2+ ions on their active sites

where the ZnO nucleation and crystal growth can be occurred. The resulted crystals will have

unidirectional nanostructure.

The effect of different molecular weights of PEG on the properties of synthesized nanotubes

was investigated while the reaction temperature and time were set to 78◦C and 24h,

respectively.

The XRD results (not shown here) indicate that regardless of PEG molecular weight, all the

products are hexagonal wurtzite ZnO with no peaks of impurities.

The morphologies of the synthesized ZnO with different PEGs (1000, 4000, and 10000) are

shown in the SEM images of Fig. 4a, 4b-c, and 4d, respectively. These figures indicate a

tubular structure in all the products. The open ends of tubes can be observed clearly in the

12
subset images. However, the diameters of the synthesized nanotubes are 600, 360, and 450

nm, which do not show a logical relationship with the molecular weight of the template. The

magnified SEM image of Fig. 4c (PEG 4000) shows a volcano-like morphology for the ZnO

structure. Such a structure may be formed since adjacent tubes are fused together to minimize

the surface energy. The primary tubes grow and weld together along the special crystal facet

[18]. This structure results in decrease in the surface area of the synthesized nanostructure.

The specific surface areas for the syntheses with PEG 1000, 2000, 4000, and 10000 are 2.49,

4.12, 2.02, and 2.51m2.g-1, respectively. The BET results confirm that the surface area of the

nanotubes synthesized by PEG4000 is lower than that of the other nanotubes because of its

special structure. The analyzing of data for surface area characterization indicates that all the

samples show type III isotherms for N2 adsorption–desorption. The interaction between ZnO

nanotubes and adsorbate are relatively week and the samples only have macropores (i.e.

multilayer adsorption process). There are no pores with an average diameter less than 50 nm,

which is confirmed by SEM images [45].

The rate constants of photodegradation of AB9 by nanotubes synthesized with PEGs having

molecular weights of 1000, 2000, 4000, and 10000 are 0.031, 0.043, 0.025, and 0.035 min-1,

respectively. The results show that the highest and lowest photocatalytic activities belong to

the synthesized nanotubes by PEG 2000 and PEG 4000, respectively. The large differences in

photocatalytic activities are related to the difference in the specific surface area of the

nanotubes.

Fig. 5 shows a typical TEM of image horn-shaped ZnO nanotubes. The contrast in the

brightness between the outer and inner sides along the length determines a hollow internal

core. The dimension of hollow internal core decreases gradually along the length and ends in

a closed tip, which is in good agreement with the SEM observations.

13
As seen in Fig. 6, the minimum value on the conductometric curve and the inflection

point in the curve of pH derivative are coincided at pHPZC=9. The surfaces of colloidal

particles in aqueous solutions are charged owing to the dissociation of surface functional

groups and preferential adsorption of ions. The surface charge of ZnO nanotubes depends on

the pH, type of ions, and concentration of the electrolyte. The point of zero charge (PZC) is

defined as the pH at which the surface exhibits a net surface charge of zero. At pH values

below the PZC, the surface is protonated and has a positive net charge while the net surface

charge turns to negative at pH values above the PZC. At pHPZC, the zero charge of the surface

allows the particles to be aggregated.

The LPS results show that the slurry with pH=7 has the narrowest size distribution.

Fig. 7 shows the histogram of size distribution for the ZnO nanotubes. The particle size

distribution histogram exhibits a narrow Gaussian distribution in the range of 1-5μm with a

peak around 2.5μm, which agrees with the results obtained from SEM images for these

nanotubes. Since the photocatalytic activity of synthesized ZnO in suspended condition will

be enhanced by increasing surface area, all the photodegradation tests were carried out in

pH=7, in which the minimum aggregation is observed.

The TGA result (Fig. 8) shows that the total weight loss of nanotubes by heating is about

5wt% mainly correspond to the losses below 350◦C. The first weight loss up to 175◦C is due

to desorption of physically adsorbed water. The second one in the range of 250–315◦C is due

to decomposition of PEG whose corresponded endothermic peak in the DSC curve is

observed at 263◦C. Since no significant weight loss is observed at higher temperatures, the

exothermic peak in the DSC curve at 495◦C may be related to the phase transition process of

the zinc oxide [46].

14
3.1.4. Effect of annealing

Annealing can be used for modifying the surface condition and crystallinity of ZnO

nanotubes. Thermal annealing also improves the crystallinity of ZnO nanotubes by

decreasing the oxygen vacancy concentration and the surface defects [47, 48]. Annealing of

the ZnO nanotubes was carried out in an air atmosphere and different temperatures (300, 350

and 400◦C) for 2.5 h. Fig. 1b shows XRD patterns of the un-annealed and annealed

nanotubes. The figure shows that the intensity of the XRD index peak is reduced with

increasing annealing temperature. The decrease in the peak intensities may be related to

reduction in the size of crystals by heating [49]. This can be confirmed by calculating the

sizes of the crystals using Schrrer equation (48 nm for un-annealed and 30 nm for annealed in

350◦C).

Fig. 9 compares the room-temperature photoluminescence (PL) spectra of un-

annealed and annealed ZnO nanotubes (T=350◦C). Photoluminescence spectra of ZnO usually

have two obvious emission peaks; the UV emission, which is related to the excitonic

recombination due to the near band-edge emission, and the visible emission, which is

commonly corresponded to the recombination of photo-generated holes with individually

ionized charge in intrinsic defects such as oxygen vacancies, Zn interstitials, or impurities.

The PL spectra of the nanotubes show a UV emission peaking at 387 nm for the annealed

sample and a wide visible emission peaking at 564 and 570 nm for the annealed and un-

annealed nanotubes, respectively. The UV emission peak intensities are increased by

annealing possibly due to decrease in the size of crystal by annealing. The peak intensity in

the visible emission region is decreased sharply by annealing. As in high temperature region

and in the presence of oxygen, re-crystallization process can be facilitated, the Zn and O

atoms will locate into their sites in the crystalline structure to reduce the Gibbs free energy to

15
a minimum stable level [50]. Therefore, the annealed ZnO nanotubes may have lower oxygen

defects and thus a lower emission peak [51]. The maximum of visible peak has a blue shift by

annealing due to decrease in the crystal size.

The photocatalytic activities of annealed are steadily increased by increasing

annealing temperature up to 350◦C. Fig. 10 indicates the rate constants of photocatalytic

reaction for un-annealed and annealed nanotubes at different temperatures. The rate constants

for the un-annealed and annealed nanotubes at, 300, 350, and 400◦C are 0.043 0, 0.045,

0.057, and 0.032 (min-1), respectively. The enhancement in photocatalytic performance of the

nanotubes by annealing in the temperature range of 300 to 350◦C can be related to increase in

their surface areas due to decrease in the size of crystals as the BET results show about 30

percent increase in the surface area by annealing at 350◦C. The figure shows that further

increase in the temperature (up to 400◦C) causes decreasing in photocatalytic activity possibly

due to agglomerating the nanotubes and decrease in their surface area [6].

3.2. Reproducibility of synthesis method

In order to confirm reproducibility of the synthesized nanotubes by the proposed

process, the synthesis in 16 h at 78◦C using PEG 2000 was repeated several times. The XRD

patterns of the synthesized nanotubes in three different trials (not shown here) represent a

hexagonal wurtzite structure with similar peak intensities for all the synthesized ZnO. The

photocatalytic performance tests were then carried out using the synthesized nanotubes. The

experiments indicate that rate constants of photodegradation reaction for all the synthesized

nanotubes are in the same range of 0.053 ± 0.0005 min-1. This confirms that the ZnO

nanotubes are well reproducible by the proposed method.

3.3. Stability of photocatalytic activity of synthesized ZnO nanotubes


16
The stability of photocatalytic activity for a photocatalyst is of high importance. In

order to evaluate the stability of synthesized nanotubes, the experiments were carried out, in

which the setup was reloaded by adding the polluting dye to make up the concentration to its

initial concentration. As shown in Fig. 11, the decrease in photocatalytic activity after five

runs is not significant that indicates proper stability of the synthesized photocatalysts. The

photocatalytic stability of the synthesized nanotubes is compared to that of commercial ZnO

nanoparticles with specific area of about 60m2.g-1. Fig. 11 shows the photodegradation rate

constants for both nanostructures. As seen in the figure, the reduction in the rate constants

after five times reusing for the nanotubes and nanoparticles are 8.8% and 90%, respectively.

UV illumination can cause photocorrosion of ZnO, which is one of the main reasons for the

decrease of ZnO photocatalytic activity in the aqueous solutions [52]. ZnO photocorrosion

occurs as a result of reaction between photo-induced holes with surface oxygen of ZnO [53,

54]. O-terminated and Zn-terminated facets in the ZnO have higher surface energies, and

these facets are more readily exposed to the corrosion [18]. As shown in Fig.12, the ratio of

Zn-terminated/O-terminated surface areas per total surface area in the nanotubes is much

lower than this ratio in the nanoparticles. Therefore, the lower reduction in the photocatalytic

activity of the nanotubes can be related to their morphology and consequently their lower

photocorrosion.

4. Conclusions

ZnO nanotubes were successfully synthesized by a facile reproducible method. The

effects of temperature, PEG molecular weight, and reaction time on the synthesis and

photocatalytic activity of ZnO nanotubes were investigated. The photocatalytic activity of

ZnO nanotubes were tested by photodegradation of AB9 as a polluting dye in the slurry

conditions under UV light. It was found that nanotubular structure cannot be synthesized in
17
temperatures below 78◦C. Furthermore, the reaction time plays an important role on

morphology of the synthesized structures i.e. while the synthesis within 16 h leads to a

complete nanotube structure, both nanotubes and nanorods are produced within 8-16 h.

Annealing ZnO nanotubes in air at 300◦C, 350◦C, and 400◦C shows that the photocatalytic

activity of the nanotubes is enhanced by increasing annealing temperature up to 350◦C while

it decreases by further increase in annealing temperature due to decrease in surface area of

the photocatalyst. The photocatalytic activity of the synthesized nanotubes demonstrates a

stable performance compared to that of commercial nanoparticles possibly due to their lower

photocorrosion.

18
Accepted Manuscript

Title: ZnO nanotubes: Preparation and photocatalytic


performance evaluation

Authors: Parvin Samadipakchin, Hamid Reza Mortaheb,


Alireza Zolfaghari

PII: S1010-6030(15)30196-9
DOI: http://dx.doi.org/doi:10.1016/j.jphotochem.2017.01.018
Reference: JPC 10506

To appear in: Journal of Photochemistry and Photobiology A: Chemistry

Received date: 18-10-2015


Revised date: 26-11-2016
Accepted date: 11-1-2017

Please cite this article as: Parvin Samadipakchin, Hamid Reza Mortaheb,
Alireza Zolfaghari, ZnO nanotubes: Preparation and photocatalytic performance
evaluation, Journal of Photochemistry and Photobiology A: Chemistry
http://dx.doi.org/10.1016/j.jphotochem.2017.01.018

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
23. Chevalier-César, C., M. Capochichi-Gnambodoe, and Y. Leprince-Wang, Growth mechanism
studies of ZnO nanowire arrays via hydrothermal method. Applied Physics A, 2014. 115(3): p.
953-960.
24. Caglar, Y., K. Gorgun, and S. Aksoy, Effect of deposition parameters on the structural
properties of ZnO nanopowders prepared by microwave-assisted hydrothermal synthesis.
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 2015. 138: p. 617-
622.
25. Chen, Y.-C., et al., Investigation of the Optimal Parameters in Hydrothermal Method for the
Synthesis of ZnO Nanorods. Journal of Nanomaterials, 2014. 2014.
26. Zhang, J., et al., Control of ZnO morphology via a simple solution route. Chemistry of
Materials, 2002. 14(10): p. 4172-4177.
27. Zhou, X., et al., Microscale sphere assembly of ZnO nanotubes. Materials Research Bulletin,
2008. 43(10): p. 2790-2798.
28. Liu, B. and H.C. Zeng, Direct growth of enclosed ZnO nanotubes. Nano Research, 2009. 2(3):
p. 201-209.
29. Tang, L., et al., Synthesis and characterization of ZnO nanorods by a simple single-source
hydrothermal method. Physica E: Low-dimensional Systems and Nanostructures, 2008. 40(4):
p. 924-928.
30. Song, J., S. Baek, and S. Lim, Effect of hydrothermal reaction conditions on the optical
properties of ZnO nanorods. Physica B: Condensed Matter, 2008. 403(10): p. 1960-1963.
31. Zolfaghari, A., H.R. Mortaheb, and F. Meshkini, Removal of N-methyl-2-pyrrolidone by
photocatalytic degradation in a batch reactor. Industrial & Engineering Chemistry Research,
2011. 50(16): p. 9569-9576.
32. Li, W.-J., et al., Growth mechanism and growth habit of oxide crystals. Journal of crystal
growth, 1999. 203(1): p. 186-196.
33. Sun, Y., D.J. Riley, and M.N. Ashfold, Mechanism of ZnO nanotube growth by hydrothermal
methods on ZnO film-coated Si substrates. The Journal of Physical Chemistry B, 2006.
110(31): p. 15186-15192.
34. McBride, R.A., J.M. Kelly, and D.E. McCormack, Growth of well-defined ZnO microparticles by
hydroxide ion hydrolysis of zinc salts. Journal of Materials Chemistry, 2003. 13(5): p. 1196-
1201.
35. Wang, J., et al., A two-step hydrothermally grown ZnO microtube array for CO gas sensing.
Applied Physics A, 2007. 88(4): p. 611-615.
36. Prevot, A.B., et al., Photocatalytic and photolytic transformation of chloramben in aqueous
solutions. Applied Catalysis B: Environmental, 1999. 22(2): p. 149-158.
37. Turchi, C.S. and D.F. Ollis, Photocatalytic degradation of organic water contaminants:
mechanisms involving hydroxyl radical attack. Journal of catalysis, 1990. 122(1): p. 178-192.
38. Lee, K.M., S.B. Abd Hamid, and C.W. Lai, Mechanism and Kinetics Study for Photocatalytic
Oxidation Degradation: A Case Study for Phenoxyacetic Acid Organic Pollutant. Journal of
Nanomaterials, 2015. 501: p. 940857.
39. Liqiang, J., et al., The preparation and characterization of La doped TiO 2 nanoparticles and
their photocatalytic activity. Journal of Solid State Chemistry, 2004. 177(10): p. 3375-3382.
40. Chen, D., X. Jiao, and G. Cheng, Hydrothermal synthesis of zinc oxide powders with different
morphologies. Solid State Communications, 1999. 113(6): p. 363-366.
41. Wang, H., et al., Controllable preferential-etching synthesis and photocatalytic activity of
porous ZnO nanotubes. The Journal of Physical Chemistry C, 2008. 112(31): p. 11738-11743.
42. Wilde, G., Nanostructured materials. Vol. 1. 2009: Elsevier.
43. Kang, Z., et al., Convenient Controllable Synthesis of Inorganic 1D Nanocrystals and 3D
High‐Ordered Microtubes. European Journal of Inorganic Chemistry, 2003. 2003(2): p. 370-
376.

20
44. Ariga, K. and T. Kunitake, Supramolecular chemistry-fundamentals and applications:
advanced textbook. 2006: Springer Science & Business Media.
45. Thommes, M., et al., Physisorption of gases, with special reference to the evaluation of
surface area and pore size distribution (IUPAC Technical Report). Pure and Applied
Chemistry, 2015. 87(9-10): p. 1051-1069.
46. Xu, Y., et al., Enhanced photocatalytic activity of new photocatalyst Ag/AgCl/ZnO. Journal of
Alloys and Compounds, 2011. 509(7): p. 3286-3292.
47. Yang, L.-L., et al., Annealing effects on optical properties of low temperature grown ZnO
nanorod arrays. Journal of Applied Physics, 2009. 105(5): p. 053503.
48. Öztürk, S., et al., Hydrogen sensing properties of ZnO nanorods: Effects of annealing,
temperature and electrode structure. International Journal of Hydrogen Energy, 2014.
39(10): p. 5194-5201.
49. Woolfson, M.M., An introduction to X-ray crystallography. 1997: Cambridge University Press.
50. Chung, J., J. Lee, and S. Lim, Annealing effects of ZnO nanorods on dye-sensitized solar cell
efficiency. Physica B: Condensed Matter, 2010. 405(11): p. 2593-2598.
51. Zhao, X., et al., Effects of thermal annealing temperature and duration on hydrothermally
grown ZnO nanorod arrays. Applied Surface Science, 2009. 255(11): p. 5861-5865.
52. Van Dijken, A., et al., Size-selective photoetching of nanocrystalline semiconductor particles.
Chemistry of materials, 1998. 10(11): p. 3513-3522.
53. Han, C., et al., Improving the photocatalytic activity and anti-photocorrosion of
semiconductor ZnO by coupling with versatile carbon. Physical Chemistry Chemical Physics,
2014. 16(32): p. 16891-16903.
54. Weng, B., et al., Toward the enhanced photoactivity and photostability of ZnO nanospheres
via intimate surface coating with reduced graphene oxide. Journal of Materials Chemistry A,
2014. 2(24): p. 9380-9389.

21
List of Tables

Table 1. Experimental conditions for synthesis of ZnO nanotube

List of Figures

Fig. 1 XRD pattern of ZnO nanotubes (a) synthesized at different reaction temperature, and

(b) un-annealed and annealed nanotubes

Fig. 2 SEM image of ZnO nanotubes synthesized at t=24h, PEG2000 and (a) 75◦C, (b) 78◦C,

(c) 80◦C, and at T=78◦C, PEG2000 and (d) 8 h, (e) 12 h, and (f) 16 h

Fig. 3 (a) Degradation of AB9 by nanotubes synthesized at 78 and 80◦C, (b) Plot of ln(C/C0)

versus time for nanotubes synthesized at 78 and 80◦C

Fig. 4 SEM images of the as-prepared ZnO products using: (a) PEG1000, (b and c) PEG4000,

and (d) PEG10000 as surfactant

Fig. 5 TEM image of nanotubes synthesized at 78◦C by PEG2000 with reaction time of 24 h

Fig. 6 Diagram of simultaneous pH metric and conductometric titration

Fig. 7 Size distribution of ZnO nanotubes synthesized at 78◦C by PEG2000 with reaction time

of 24 h dispersed in pH = 7

Fig. 8 TGA and DSC of ZnO nanotubes synthesized at 78◦C by PEG 2000 with reaction time

of 24 h

Fig. 9 PL spectra of un-annealed and annealed nanotubes

Fig.10. Degradation of AB9 by un-annealed and annealed nanotubes at 300, 350, and 400°C

Fig. 11. Histogram of K versus frequent use of ZnO nanotubes and commercial ZnO

nanoparticles under UV-light irradiation for five consecutive cycles

Fig. 12 Comparison of photocorrosion characteristics of ZnO (a) nanotube and (b)

nanoparticle.

22
Table 1.

Experimental conditions for synthesis of ZnO nanotube

Sample T Mw of PEG Reaction time ZnO


Morphology
No. (◦C) (g/mol) (h) formation
1 70 2000 24 No No
2 75 2000 24 No No
3 78 2000 24 Yes Yes
4 80 2000 24 Yes Yes
5 78 1000 24 Yes Yes
6 78 4000 24 Yes Yes
7 78 10000 24 Yes Yes
8 78 2000 16 Yes Yes
9 78 2000 12 Yes No
10 78 2000 8 Yes No

23
a Zn(OH)2 b o 700
400 C
8000 850
70oC o
350 C

Intensity (Counts)
75oC
Intensity (Counts)

870
o
002 101 300 C
100

2500

112
110

103
4000

201
102

200
78oC

unannealed
80oC
0

0 30 60 0 30 60
2 (degree) 2 (degree)

Fig. 1 XRD pattern of ZnO nanotubes (a) synthesized at different reaction temperature, and (b) un-

annealed and annealed nanotubes

24
T

10 µm 10 µm 10 µm

5 µm 5 µm 5 µm
t

Fig. 2 SEM image of ZnO nanotubes synthesized at t=24h, PEG 2000 and (a) 75°C, (b) 78°C,

(c) 80°C, and at T=78°C, PEG 2000 and (d) 8 h, (e) 12 h, and (f) 16 h

25
T= 80oC
100 a T= 78oC

80

60
% (C/C0)
40

20

0 20 40 60 80 100 120
Time (min)

b T=78oC
0
T=80oC

-1 k = 0.0221

-2
Ln(C/C )
0

k = 0.0433
-3

-4

-5

-6
0 20 40 60 80 100 120
Time (min)

Fig. 3 (a) Degradation of AB9 by nanotubes synthesized at 78 and 80°C, (b) Plot of ln(C/C0)

versus time for nanotubes synthesized at 78 and 80°C

26
Fig. 4 SEM images of the as-prepared ZnO products: (a) using PEG 1000 (b and c) using

PEG 4000 (d) using PEG 10000 as surfactant

27
Fig. 5 TEM image of nanotubes synthesized at 78°C by PEG 2000 with reaction time of 24 h

28
320 0.005
300 0.000
280
-0.005
Conductivity (S/cm)

260

Derivative pH
240 -0.010

220 -0.015
200
-0.020
180
-0.025
160
Conductivity
140 -0.030
Derivative pH
0 200 400 600 800 1000 1200
V(l)

Fig. 6 Diagram of simultaneous pH metric and conductometric titration

29
100

80

60
Intensity

40

20

0
0.01 0.1 1 10 100
particle size (m)

Fig. 7 Size distribution of ZnO nanotubes synthesized at 78°C by PEG 2000 with reaction

time of 24 h dispersed in pH = 7

30
100
exo 0.15

99
0.10

DSC (mW)
98
TG(%)

0.05

97
0.00

96 -0.05

0 100 200 300 400 500 600 700


o
Temperature ( C)

Fig. 8 TGA and DSC of ZnO nanotubes synthesized at 78°C by PEG 2000 with reaction time

of 24 h

31
80 annealed at 350oC
70

60 unannealed
Intensity (a.u)
50

40

30

20

10

0
300 350 400 450 500 550 600 650

(nm)

Fig. 9 PL spectra of un-annealed and annealed nanotubes

32
Un-annealed
100 o
300 C
o
350 C
o
80 400 C

60
%(C/C0)

40

20

0 20 40 60 80 100 120

Time (min)

Fig.10. Degradation of AB9 by un-annealed and annealed nanotubes at 300, 350, and 400°C

33
0.10
ZnO nanotubes
0.09 ZnO nanoparticles

0.08

0.07

) 0.06
-1
K(min

0.05

0.04

0.03

0.02

0.01

0.00
1 st 2 nd 3 rd 4 th 5 th

Number of runs

Fig. 11. Histogram of K versus frequent use of ZnO nanotubes and commercial ZnO

nanoparticles under UV-light irradiation for five consecutive cycles

34
Fig. 12 Comparison of photocorrosion characteristics of ZnO (a) nanotube and (b)

nanoparticle.

35

You might also like