Moumita Pal,
Susanta Bera,
Saswati Sarkar and
Sunirmal Jana*
Sol–Gel Division, CSIR-Central Glass and Ceramic Research Institute (CSIR-CGCRI), P.O. Jadavpur University, 196 Raja S. C. Mullick Road, Kolkata-700032, West Bengal, India. E-mail: sjana@cgcri.res.in; janasunirmal@hotmail.com; Fax: +91-33-2473 0957; Tel: +91-33-2483 8082 Tel: +91-9432355818 (M)
First published on 12th February 2014
Al doped nanostructured zinc oxide thin films (thickness, 165 ± 5 nm) on silica glass were prepared from zinc acetate based solutions of varying dopant content (0 to 10 at% with respect to Zn). X-ray diffraction confirmed the presence of nanocrystalline hexagonal ZnO. On increasing the doping level, we observed a change in ZnO morphology (spherical, hexagon and plate-shaped) under field emission scanning and transmission electron microscopes and a gradual decrease of ZnO crystallite size (14.0 to ∼10 nm) vis-à-vis an enhancement of direct band gap energy of the films. Root mean square film surface roughness and chemical state of elements were studied by atomic force microscopy and X-ray photoelectron spectroscopy respectively. In addition to common intrinsic defects in ZnO, a defect (designated as D˙ZA, appearing as a paramagnetic singly negatively charged oxygen vacancy) was identified up to 4% doping from the appearance of photoluminescence emission at 398 nm and measurement of paramagnetic property of the films. Film photocatalytic activity towards Rhodamine 6G dye decomposition was performed under UV (254 nm) irradiation and the film with 4% doping (A4ZO) showed the highest value of first order decomposition rate constant. On increasing Al content, the trend of change of defect concentration (oxygen vacancies) analyzed by Raman spectra was found to be identical with the dye photodecomposition activity (PA) of the films. A major role of DZA˙ towards the PA was explored under visible light. We proposed the reaction mechanism of PA based on the experimental results. The A4ZO would be expected to decompose micro-organisms under visible light.
The widespread applications of ZnO primarily based on its exceptional properties5,6 such as microstructural, electrical, optical, electronic, optoelectronic. Moreover, by doping with the elements belonging to group III (B, Al, etc) of periodic table, optical and photocatalytic properties can significantly be changed depending7–11 upon the nature and content of the element. This is because incorporation of doping element into the ZnO crystal, the lattice parameters and defects (intrinsic/extrinsic) concentration7–11 would change. The content of doping also depends upon the crystal size due to self purification12 in semiconductor nanocrystals. Depending upon the preparation method13 as well as doping,14 many defects13 have been identified in ZnO both theoretically and experimentally. Formation of the defects would occur within the band gap of ZnO (bulk direct band gap, 3.37 eV).15,16 These defects are generally characterized by photoluminescence studies.11,17 The nature and concentration of defects18 also influence the magnetic property of ZnO. Intentional creation of defects has also their own perspectives (usefulness).10,13,14,19 For example, it is generally accepted that oxygen vacancies function as electron acceptors19 which can trap the photogenerated electrons and therefore, the efficiency of the catalyst could be enhanced. In addition, the catalytic application can also be extended under visible light irradiation in presence of the defects.
Various defects including the oxygen vacancy can play an important role15,20–26 on the photocatalytic activity (PA) of ZnO. Further, PA also found to depend on concentration of oxygen vacancy. Sun et al.26 observed a higher photocatalytic activity due to the existence of high-level surface and subsurface oxygen vacancy in well-aligned single-crystalline ZnO nanobelt arrays on Si wafer. Also, Wang et al.22 reported an enhanced photocatalytic activity of ZnO nanoflowers due to the presence of large extent of oxygen vacancy which can able to separate the photogenerated electron–hole pairs.
Several ZnO film deposition techniques such as sputtering (e.g. rf sputtering, magnetron sputtering), CVD/PVD, sol–gel etc. are known. However, sol–gel is a facile and cost effective technique for tailoring film crystallinity, morphology and optical properties through several competing factors6,9 such as sol/solution chemistry, dopant content and nature, curing condition, etc. Many authors already also reported7,10,11,27 the morphology, optical, opto-electronic and electrical properties of aluminium zinc oxide (AZO) films. The film can also act a photocatalyst towards photodegradation of organic dyes but a very few reports11,28 are available in this concern. Very recently, Pradhan et al.11 reported the photocatalytic activity towards degradation of methyl orange on exposure of different light sources (UV, 380 nm; white light and sun light) of AZO thin film deposited by spray pyrolysis. They claimed a high rate of degradation (100% under UV and sunlight in 3 h irradiation) using 25% Al doped ZnO film as photocatalyst.
It is also known that the rate of decomposition29 of an organic dye depends upon its chemical structure30 and optimum amount of catalyst.31 Moreover, a photocatalyst would not give similar nature of decomposition for all the dyes even the measurement conditions remain same. In addition, most of the photocatalytic activities of powders15,19 have been reported on azo dyes.32–34 This is probably, 70% azo dyes use in textile industry35 and remaining 30% dyes including rhodamine 6G (Rh-6G)36 are mostly used as laser dyes. However, the thin film photocatalyst may find application in the area6 of human health and environment. To the best of our knowledge, the decomposition of Rh-6G dye using sol–gel based Al-doped ZnO thin films as photocatalyst seems not found in the literature.
Hence, in the present work, a systematic study was performed on Al doped nanostructured zinc oxide thin films derived from zinc acetate based precursor solutions of varying dopant content (0 to 10 at% with respect to Zn) to observe the effect of doping on crystallinity and lattice parameters, morphology and microstructure, optical (band gap and photoluminescence), magnetic and photocatalytic properties of the films. Photocatalytic activity of the films towards Rh-6G dye decomposition was studied under UV (254 nm) irradiation. In addition to common intrinsic defects of ZnO, a defect (designated as DZA˙, appeared as paramagnetic singly negatively charged oxygen vacancy) was identified up to 4% doping from the measurement of photoluminescence and paramagnetic property of the films. Moreover, on increasing doping level, the trend of change of defect concentration (oxygen vacancies) analyzed by Raman spectra was found identical with the dye photodecomposition activity of the films. Further, a major role of the D˙ZA defect towards photocatalytic activity was explored under visible light.
Thin films were deposited on pre-cleaned9 glass substrate by dipping technique (Dip Master-200, Chemat Technology Inc., USA) with a constant withdrawal speed of 12 cm min−1 utilizing the aged solutions. To get the oxide films, the as-coated samples were initially dried in an air oven at 60 °C followed by curing at 500 °C with 1 h soaking time under air atmosphere in an electrical furnace. The final films derived from the aged solutions, SA0ZO, SA1ZO, SA2ZO, SA3ZO, SA3.5ZO, SA4ZO, SA4.5ZO, SA6ZO and SA10ZO were designated as A0ZO, A1ZO, A2ZO, A3ZO, A3.5ZO, A4ZO, A4.5ZO, A6ZO and A10ZO respectively.
Photocatalytic activity (PA) of the films towards degradation of Rhodamine 6G (Rh-6G) was studied in a custom built stainless steel UV (wavelength, 254 nm) curing chamber. The coated sample (total area, ∼3.5 cm2) was placed approximately 45° angle with the wall of borosilicate glass beaker of 100 ml capacity containing the solution of dye (10−5 M, Co) in double distilled water. The detailed PA measurement set up has already been reported elsewhere.6 We also carried out the PA of a specific film (A4ZO) under visible light irradiation. For this purpose, a tungsten bulb (200 W, Philips India) was used as a visible light source employing two cutoff glass filters (cutoff wavelengths, 393.5 nm and 412 nm), the details of the setup is given under the results and discussion (Fig. S4, ESI†). In a certain time interval, ∼4 ml dye solution took out and visible absorption spectrum was recorded to find out the remnant concentration (C) of the dye with the help of a calibration curve of the dye solutions. The calibration curve was constructed by plotting dye concentration against absorbance (OD) at 527 nm peak wavelength of Rh-6G solutions obeying Lambert–Bayer's law. The PA of the films was analyzed by plotting ln(Co/C) (dye concentration; Co, initial and C, remnant) versus irradiation time. The rate constants of decomposition reaction (considering first order reaction kinetics) were calculated from the plots.
Different surface feature (Fig. 2) of the films was visualized from FESEM images. A change in shape and size of ZnO clusters (spherical, near hexagon and plate-like) was observed in different doping levels. On increasing Al content, the shape of ZnO clusters found from spherical (undoped, size range: 66–83 nm) to near hexagon (1% doped, size range of largest side: 499–832 nm) to plate-like (2–4% doped, average thickness: 60–80 nm) and finally to nearly spherical (≥6% doped, size range: 60–70 nm). The hexagon-like clusters were supposed to be floated on the bed of spherical particles while plate-like clusters were found to be embedded into the featureless matrix. The FESEM images of the films supported that the spherical clusters would be the primary one from which different shapes of ZnO clusters originated39 through unequal growth rate of polar and non-polar surfaces of hexagonal ZnO crystals by Ostwald ripening. This would happen39 due to different chloride (Cl−) concentrations coming from aluminium chloride used in the precursor solutions.
A change of ZnO film surface morphology was observed by FESEM (Fig. 2) study. To understand the variation of film surface roughness due to the change of surface morphology, AFM measurement was carried out. A plot of root mean square surface roughness (RMSSR) versus dopant content is shown in Fig. S6 (ESI†) where the lowest and highest roughness value was found in 6% and 2% doped films respectively. In fact, in other films including 4% doped film (A4ZO), an approximately same roughness value was obtained. The plot also displays individual film 3D surface topography (insets, Fig. S6†).
TEM image (Fig. 3a) of undoped film (A0ZO) shows agglomerated quasi-spherical ZnO nanoclusters. However, the agglomeration found to be diminished (Fig. 3b, c, f and g) in doped films. We calculated the average size of nanoclusters from the TEM images (histograms, insets of Fig. 3a and b for A0ZO and A1ZO respectively; Fig. 3d for A4ZO; HRTEM images in the inset of Fig. 3f and inset m of Fig. 3g for A6ZO and A10ZO respectively). The sizes were 21.7, 20.8, 16.8, 14.7 and 11 nm for A0ZO, A1ZO, A4ZO, A6ZO and A10ZO films respectively. The trend in change of the nanocluster size as a function of doping content also supported the XRD result (Fig. 1b). However, the size measured from TEM images was found always higher than the calculated size from the XRD reflections. This might be considered due to agglomeration42 towards clustered formation. A closed porous structure was found distinctly in the TEM image (Fig. 3b) of A1ZO film. Content (at%) of Zn and Al present in the films were analyzed by TEM-EDS analysis and found approximately same to the content used in the precursor solutions. As for example, the content of Zn (Zn-k) and Al (Al-k) in A4ZO film (Fig. 3e for A4ZO) was 95.0 ± 1.0 and 4.8 ± 0.8% respectively whereas in A10ZO film (inset n of Fig. 3g), these were 90.1 ± 1.0 and 9.90 ± 0.9% respectively.
Fig. 4 shows the substrate corrected FTIR spectra of the films. In all the films, a strong vibration appeared at 414 cm−1 indicated the presence of h-ZnO43 which supported the XRD result (Fig. 1a). Moreover, the spectral region, 550–450 cm−1 of Zn–O stretching vibration9 found to broaden with respect to the undoped film. This would be due to incorporation of Al into the ZnO crystal lattice. Although, no clear FTIR vibration for Al–O was found up to 6% doped films but a weak and broad vibration appeared at ∼805 cm−1 in A10ZO. This might be an indication for the presence of Al–O44 in the film network.
It is known that absorption co-efficient (α) and optical band gap energy (hν, Eg) are interrelated9 for a direct transition semiconductor, for example, ZnO. The Eg of the films was calculated from eqn (1). Individual α value was determined using respective UV-visible absorption and film physical thickness. Finally, the plot of (αhν)2 versus hν was drawn (Fig. 5a) and observed that the plots are parabolic in nature with two distinct inflexions indicated two transitions occurring in ZnO film. At the inflexion, the tangential extrapolation of the linear part of the plot to X-axis, (αhν)2 = 0 gives the Eg value. Among the inflexions, one is located at ∼3.3 eV for all the films. This would correspond9 to the bulk Eg of ZnO. Another Eg (inset, Fig. 5a) was observed at 3.47, 3.57, 3.62, 3.67, 3.70, 3.81 and 3.88 eV for 0, 1, 2, 3, 4, 6 and 10% doped films respectively. This could be related due to the effect of size9 of nanostructured ZnO.
αhν = A(hν − Eg)1/2 | (1) |
All the films showed UV-Vis photoluminescence (PL) emissions (Fig. 5b) when excited at 340 nm (Inset, Fig. 5b). Undoped ZnO film (A0ZO) showed an intense emission at ∼424 nm in addition to the emissions at ∼398 nm (broad shoulder), ∼450 nm, ∼485 nm and ∼530 nm. The intensity of emission at 398 nm (Table S1, ESI†) with respect to the other emissions found to gradually increase with doping and reached a maximum value in 4% doped film. Although, the emission at 398 nm found to be nearly disappeared in ≥6% doped films but a new UV PL emission at 376 nm (∼3.30 eV) was appeared. However, except the emission at 376 nm, the intensity of the emissions appeared above 400 nm found to decrease in ≥6% doped films. In addition, a shifting of the emission at ∼450 nm to ∼456 nm was noticed in ≥6% doped films. The UV PL emission at 376 nm corresponded to band to band transition11 through the recombination of electrons with holes in the valence band as free exciton recombination whereas the emission at 398 nm would consider as crystal defect (designated as DZA˙, appeared as singly negatively charged oxygen vacancy like , discussed later). Interestingly, in ≥6% doped films; the intensity of emission at 376 nm was always higher than all the visible emissions. It is known that the emission at 376 nm, a characteristic UV emission45 that could generally be appeared in low concentration of crystal defects in ZnO. Therefore, it would suggest11 that there would be a very low concentration of defects including the DZA˙ in ≥6% doped (A6ZO and A10ZO) films. Previously, some researchers also reported the UV PL emission including the peak at 398 nm as band edge emission45–47 of nanostructured ZnO. Very recently, an important remark on the 398 nm emission in Sn doped ZnO film has been given by Shi et al.48 They observed that the intensity of the emission increased with doping level and described the defect as Sn related doubly ionized donor formed in the deep states of ZnO band gap. Also, in Y doped ZnO nanoparticles, the same emission was reported.45 In fact, Yu et al.49 noticed the PL emission at 398 nm along with 378 and 420 nm in indium oxide nanofibers and demonstrated the emissions originating from different energy levels of oxygen vacancies. However, in the present work, we observed the emission in undoped ZnO film and the intensity of the emission increased with doping level of up to 4%. Therefore, Al doping could tailor the concentration of the defect within the band gap of ZnO (scheme, Fig. 10b). However, the defect already originated during the film deposition process. The defect would function as a singly ionized donor similar to the defect in Sn doped ZnO (doubly ionized donor)48 and would behave as hole scavengers in the Al doped ZnO. However, it could not be understood for the disappearance of the emission in ≥6% doped films.
Generally, five intrinsic defects (oxygen and zinc interstitials, oxygen and zinc vacancies and antisite oxygen)13 could be formed within the band gap of ZnO. These defects could able to generate PL emissions after excitation with photons. However, the formation of the defects was dependent on preparative methods13,45 and doping.14 In this work (Fig. 5b), single ionized (singly positively charged) oxygen vacancy () was characterized from the appearance of PL emission at ∼450 nm (ref. 50) and its red shifting might be related to the change in defect concentration.45 Moreover, the presence of zinc interstitials (Zni)13 was also supported from the PL emission at 424 nm. Further, the existence of neutral oxygen vacancy (
) would be ascertained from the observation of emission at 530 nm.51,52 Also, the emission at 485 nm supported50,53 the formation of antisite oxygen. Perhaps, singly positively charged oxygen vacancy (
) and/or neutral oxygen vacancy (
) would be the source54,55 of the DZA˙ defect. Anyhow, the enhancement of DZA˙ defect concentration (Table S1, ESI†) would proceed via a process of accepting electron/(s) by
and/or
charge state/s from the conduction band of nanostructured ZnO crystals. Consequently, the process would be accelerated on increasing the number of conduction electrons which believed to be a maximum at the Al doping level of 4%.
Presence of various defects (DZA˙, zinc interstitials, antisite oxygen, neutral oxygen vacancy, singly positively charged oxygen vacancy) was identified by the PL spectra (Fig. 5b) of the films. It was also found that the relative concentration of DZA˙ (calculated from the PL emission peak intensity) (Table S1, ESI†) found to increase on doping level (up to 4%). ZnO is known18 to be a diluted magnetic semiconductor (DMS). One can tailor the nature and concentration of defects18 as well as the magnetic property through suitable doping. Hence, measurement of magnetic property (Fig. 6) was done on three representative samples (A0ZO, A4ZO and A10ZO) to realize the effect of the defects, especially the defects of paramagnetic in nature on the magnetic property of the films. Generally, five charged states of oxygen vacancies,55 termed as ,
,
,
and
would possible to form in oxides. These charge states correspond to zero, one, two, three and four electrons in the vicinity of the oxygen vacancies respectively. Therefore, the only
, and
oxygen vacancies would be considered as paramagnetic states due to the presence of unpaired electrons. It was also reported that zinc interstial as identified from PL spectra (Fig. 5b) in the present work was diamagnetic56 owing to the electron configuration ending with 4s2. The presence of antisite oxygen (as characterized by PL) in 2-charged state like lattice oxygen (O2−) and/or interstitials oxygen (O2−) in h-ZnO would be diamagnetic.56 However, some oxygen vacancies54–57 (singly positively charged,
and singly negatively charged,
) would be paramagnetic. It should be noted that the
state was recognized from the observation of PL emission at 450 nm (Fig. 5b) while the DZA˙ defect was identified from the emission at 398 nm. We considered, the DZA˙ defect as singly negatively charged oxygen vacancy, like
. A justification on the proposition could be given on the basis of magnetic property measurement. At 15 kOe applied magnetic field, the A4ZO film, showed highest paramagnetic contribution in terms of magnetization value, 16.79 ± 1.13 emu cm−3 (Fig. 6b) whereas in the A0ZO and A10ZO films, the values were 2.62 ± 0.11 (Fig. 6a) emu cm−3 and 2.33 ± 0.24 emu cm−3 (with weak ferromagnetic ordering) (Fig. 6c) respectively. Thus, the magnetization value obtained in A4ZO is more than 6 times higher than the other films. This could strongly support the enhancement of paramagnetic defect concentration in 4% doped film (A4ZO). As the PL study (Fig. 5b and Table S1, ESI†) already evidenced the increase in concentration of DZA˙ with respect to
, hence, the highest paramagnetic property could be due to the highest concentration of DZA˙ defect in A4ZO. On the other hand, a weak ferromagnetic ordering (Fig. 6c) in absence of DZA˙ observed in A10ZO would be due to the presence of other defects.58 Probably, the paramagnetic DZA˙ defect would be responsible for preventing the ferromagnetic ordering in A0ZO and A4ZO ZnO films.
![]() | ||
Fig. 6 Substrate corrected magnetization versus applied magnetic field curves of undoped and doped ZnO films. |
XPS study of three representative samples, undoped (A0ZO), 4% doped (A4ZO) and 10% doped (A10ZO) films were performed (Fig. 7 and S1, ESI†). The position of binding energy signals (Fig. 7a) for Zn2p1/2and Zn2p3/2 core levels in A0ZO, A4ZO and A10ZO films were observed at 1044.65 ± 0.05 eV and 1021.9 ± 0.1 eV respectively. The difference between the two peaks is 23.1 ± 0.1 eV, matched well with the reported48 value for Zn(II). A broad signal of O 1s (Fig. 7b and c and S1b, ESI†) was observed in both the films but the signal for A0ZO film was more asymmetric in nature than A4ZO and A10ZO films. Moreover, the broad signals could be resolved into three prominent Gaussian peaks centred at 530.2 ± 0.1 eV (S1), 531 ± 0.05 eV (S2) and 532.2 ± 0.1 eV (S3) for A0ZO film (Fig. 7b); 530.5 ± 0.2 eV (S1), 531.2 ± 0.1 eV (S2) and 532.2 ± 0.1 eV (S3) for A4ZO film (Fig. 7c) and 530.2 ± 0.1 eV (S1), 531.2 ± 0.1 eV (S2) and 532.2 ± 0.1 eV (S3) for A10ZO film (Fig. S1b, ESI†). The appearance of S1 was due to O2− ions59 of wurtzite hexagonal ZnO as evident from XRD (Fig. 1a) while S2 could relate to O2− ions of oxygen deficient ZnO. Further, the S3 peak ascribed to the existence of loosely bound oxygen from carbonate, adsorbed water or oxygen molecules on the film surface. The S1, S2 and S3 peaks matched well with the reported values.59 On the other hand, the maximum concentration of O2− signal (S2) at ∼531 eV (due to oxygen deficiency) with respect to total oxygen signal was observed (Table S2, ESI†) in A4ZO film. This result also supported the PL data where the relative emission peak intensity at 398 nm (DZA˙ defect, appeared as singly negatively charged oxygen vacancy) increased (Table S1, ESI†) up to the doping level of 4%. The binding energy signals (Fig. 7d) for Zn3p and Al2p core levels observed prominently in A4ZO film. However, the signal for Al2p core level became very strong (Fig. S1a, ESI†) in A10ZO film. This could be due to the increase of Al doping level (2.5 times higher than A4ZO). The binding energy peak (symmetric in nature) for Al2p3/2 core level was found at 74.1 ± 0.05 eV in the doped films. The observed peak was slightly lower than the reported value of stoichiometric Al2O3, suggested the presence of oxygen deficiency in the film network.
ZnO is an n-type semiconductor oxide14 due to the presence of natural oxygen vacancies along with other common (intrinsic) defects. These could be tailored by incorporation of dopant (e.g. Y into ZnO crystal). In the periodic table of elements, ‘Al’ belongs to the same group (Group IIIA) of ‘Y’. In the present work, the XRD analysis (inset of Fig. 1b) already confirmed a maximum amount, ∼4% of Al entered into Zn site of hexagonal ZnO. Hence, the introduction of Al in ZnO crystal, there could be a change in concentration of defects depending upon the content of incorporated dopant. On the other hand, by Raman spectral study, the defect concentration (oxygen vacancies)60 in addition to the active mode of vibrations in h-ZnO could be better understood. Therefore, we performed the Raman study on the Al doped ZnO films. The equation for optical phonon vibrations of hexagonal wurtzite ZnO crystal at the Γ-point can be expressed by the irreducible representation9 as shown in eqn (2),
Γ = A1 +2B1 + E1 + 2E2, | (2) |
![]() | ||
Fig. 8 Substrate corrected Raman spectra of undoped and doped ZnO films. Inset shows a plot of Raman peaks intensity ratio of E1/E2 with dopant content. |
Photodecomposition study of Rh-6G dye (10−5 M) (Fig. 9, 10a and S3 to S5 of ESI†) using the films as photocatalyst was performed under UV (λ, 254 nm) and visible light irradiations. The photocatalytic activity for 4% doped film (A4ZO) was studied on exposure of visible light employing cut off glass filters (Filter-1: cut off wavelength, 393.5 nm and Filter-2: cut off wavelength, 412 nm). Considering the first order reaction kinetics, the dye decomposition (Fig. 9a, S3, S5a (inset) and S5, ESI†) rate constant (k) was measured (Fig. 9b, c and d and S5a, ESI†). A plot (Fig. 10a) of k value (in the order of 10−4 min−1) against doping content was drawn. From the plot, it is noticed that the ‘k’ value rapidly increased on increasing dopant content up to 4% (in A4ZO film) and then decreased on further increase of the dopant content. However, the lowest ‘k’ value was found in 10% doped (A10ZO) film would arise due to the faster rate of recombination23,62 of photogenerated electrons and holes in relatively lowest defect concentration (mainly DZA˙) of nanostructured ZnO. On the other hand, the generation of highest rate of photodecomposition (highest k value) in A4ZO would consider as the presence of highest defect concentration as confirmed from Raman spectral analysis (inset of Fig. 8a and 10a) which prevented the faster rate of recombination23,62 of photogenerated electrons and holes. As a result the enhancement of the photocatalytic activity could be occurred.
Generally, photocatalytic activity of semiconductor oxide depends upon many factors6,11,23,63 such as surface area (inversely related to size and directly proportional to the band gap energy), surface structure (related to film surface morphology), defect concentration, etc. In the present work, the A4ZO film showed the highest photocatalytic activity although the lowest size of ZnO was found in A10ZO film. It was also found that the size of ZnO gradually decreased with increasing Al content as evidenced from XRD (Fig. 1) and TEM (Fig. 3) vis-à-vis the direct band gap energy gradually increased (Fig. 5). Hence, the change of size/band gap energy of ZnO would not be a deciding factor to control the photodecomposition activity of the dye. In addition, the change of root means square surface roughness (relate to surface structure) measured from AFM for a fixed area of the films (Fig. S6, ESI†) would not support the change of the dye decomposition rate constant. This is because the lowest and highest surface roughness value were found in 6% (A6ZO) and 2% doped (A2ZO) films respectively whereas approximately same value was obtained in the other films including 4% doped (A4ZO) film. Further, it is known that the faster rate of electron–hole recombination63 could be prevented by introduction of surface defect in ZnO crystal. As a result, the enhancement of photocatalytic activity could be expected. This intentional defect creation in semiconductor photocatalysts could also be an attractive way of photocatalytic studies63 under visible light irradiation. In this work, the Raman spectral analysis showed a systematic change of defect concentration (oxygen deficiencies) with doping content (inset, Fig. 8 and 10a) and a maximum defect concentration was found in A4ZO film. On increasing dopant content, the trend (Fig. 10a) of change in defect concentration was identical to the trend of change of the dye decomposition rate constant. This result strongly suggested that the defect concentration controlled the rate of the dye decomposition. However, it was absolutely necessary to understand the specific role of the DZA˙ defect as one of the oxygen deficiencies in comparison to defects such as singly positively charged oxygen vacancy, neutral oxygen vacancy, interstitial Zn, antisite oxygen in ZnO as already evidenced from PL (Fig. 5b) and XPS (Fig. 7b and c and S1b, ESI†) spectral study particularly in A4ZO film. To find out the specific role of the DZA˙ defect towards the dye decomposition, we used cut off glass filters, Filter-1 (cut off wavelength, 393.5 nm) and Filter-2 (cut off wavelength, 412 nm) (Fig. S4b†) with the identical visible transmittance during the measurement of photocatalytic activity of the film under visible light irradiation (scheme: Fig. S4a†). Using Filter-1, we obtained the ‘k’ value (3.0 × 10−3 min−1) (inset, Fig. S5a†) slightly lower than the ‘k’ value (3.3 × 10−3 min−1 measured under UV irradiation, Fig. 9c) using A4ZO film as photocatalyst. However, a very low rate of dye decomposition (Fig. S5b†) was found when the Filter-2 was used. This result could confirm the major role of the DZA˙ defect towards the dye decomposition. The lowest rate of the dye decomposition using Filter-2, would be due to inactiveness of the DZA˙ defect under lower energy (below ∼3.0 eV, 412 nm) of irradiation than the minimum energy (∼3.12 eV, 398 nm) required for activation of the defect (DZA˙) towards photocatalytic decomposition of the dye. It is known that the oxygen vacancies as electron acceptors19 could trap the photogenerated electrons. In the present work, the DZA˙ defect would function as hole scavengers because from the magnetic moment study (Fig. 6), it appeared as paramagnetic like charge state (singly negatively charged oxygen vacancy). In addition, from the PL spectral study, it was verified that the other common defects including singly positively charged oxygen vacancy,
were present in all the films. These would also take part in the photocatalytic activity. However, our experimental results on the dye decomposition under visible light using cut off glass filters clearly demonstrated that these defects including
found to be less effective than the DZA˙ defect. Hence, a major contribution of the DZA˙ defect together with a minor role of the other defects including singly positively charged oxygen vacancy of sol–gel based nanostructured Al doped ZnO films was found in the photocatalytic decomposition of Rh-6G dye.
On the basis of our experimental results, a scheme19 for the mechanism of photocatalytic reactions using the nanostructured ZnO film as photocatalyst is given in Fig. 10b.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c3ra44612c |
This journal is © The Royal Society of Chemistry 2014 |