Next Article in Journal
The Effect of Coating Density on Functional Properties of SiNx Coated Implants
Previous Article in Journal
Fracture Characteristics and Analysis in Dissimilar Cu-Al Alloy Joints Formed via Electromagnetic Pulse Welding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Properties of CaCO3-Supported Nano-TiO2 Composite with Improved Photocatalytic Performance

1
Beijing Key Laboratory of Materials Utilization of Nonmetallic Minerals and Solid Wastes, National Laboratory of Mineral Materials, School of Materials Science and Technology, China University of Geosciences, Xueyuan Road, Haidian District, Beijing 100083, China
2
Beijing Building Materials Academy of Sciences Research Co., Ltd., Shixing Street, Shijingshan District, Beijing 100041, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(20), 3369; https://doi.org/10.3390/ma12203369
Submission received: 17 September 2019 / Revised: 9 October 2019 / Accepted: 12 October 2019 / Published: 15 October 2019

Abstract

:
In order to improve the photocatalytic degradation efficiency of nano-TiO2, reduce its usage and realize recycling and reuse, CaCO3–TiO2 composite photocatalyst was prepared with calcium carbonate (CaCO3) and TiO2 in a grinding machine through the integration of grinding depolymerization, dispersion and particle composition. The photocatalytic degradation performance, recycling performance, structure and morphology of CaCO3–TiO2 were studied. The interaction mechanism between CaCO3 and TiO2 and the improvement mechanism for the photocatalytic performance of TiO2 were also discussed. The results show that under the UV light irradiation for 20 and 40 min, the degradation efficiency of methyl orange by the composite photocatalyst with 40% TiO2 (mass fraction) was 90% and 100%, respectively. This was similar to that of pure TiO2, and the performance of the composite photocatalyst was almost unchanged after five cycles. CaCO3–TiO2 is formed by the uniform loading of nano-TiO2 particles on the CaCO3 surface, and the nano-TiO2 particles are well dispersed. Due to the facts that the dispersion of nano-TiO2 is improved in the presence of CaCO3 and the charge transport capability is improved through the interfacial chemical bonds between CaCO3 and TiO2, the formation of this complex is an intrinsic mechanism to improve the photocatalytic efficiency of nano-TiO2 and reduce its usage in application processes.

1. Introduction

In recent decades, with the rapid development of industries such as chemical industry, mining industry and agricultural breeding industry, environmental pollution problems, including water pollution, have become increasingly serious. Photocatalytic techniques are considered to be one of the most ideal ways to combat environmental pollution problems such as water pollution [1,2]. Therefore, it is of great significance to optimize the use conditions of photocatalysts, improving the function of photocatalyst technology and reducing the use amount of photocatalyst required to reduce the overall cost. Nano-TiO2 has become the most widely studied and applied semiconductor photocatalytic materials owing to its non-toxic, non-polluting and stable properties [3,4]. Under the excitation of ultraviolet light, electron–hole pairs are produced by electron transition in nano-TiO2, following which the separated electrons and holes migrate to the surface of TiO2 to produce active groups or directly react with pollutants. Thus, the degradation of pollutants such as dyes, heavy metal ions, phenols, dichlorophenol and antibiotics can be achieved through the further redox action [5,6,7,8]. However, there are still the several problems in the application of TiO2. Firstly, nano-TiO2 is easily agglomerated due to its high surface energy, which leads to the reduction of specific surface area and surface active sites, thus reducing its photocatalytic activity. Secondly, the loss of nano-TiO2 is serious when it is used in water, and is difficult to recover. The loss of nano-TiO2, caused by the lack of reuse, not only increases the use cost but also causes secondary pollution [9,10]. Many studies have shown that loading nano-TiO2 on the surface of the carrier with a large particle size to prepare a composite photocatalyst can effectively solve the above problems [11,12,13]. At present, the carriers for nano-TiO2 are mostly natural minerals and inorganic materials such as glass and ceramics. Among them, the natural mineral carriers that have been greatly studied are diatomite, zeolite, kaolin and montmorillonite [14,15,16,17]. However, these minerals generally need to be processed, such as via ore dressing and purification, to meet the requirements for catalyst carriers. This increases the cost of the composite photocatalyst, which undoubtedly restricts their industrial production and application.
Calcium carbonate (CaCO3) is an important inorganic mineral material, mainly made of non-metallic mineral raw materials such as calcite. It has the characteristics of high purity, high whiteness and low cost, and is obviously superior to most non-metallic minerals in cost-performance [18,19]. Therefore, the construction of CaCO3–nano-TiO2 composite photocatalyst using CaCO3 as a carrier can not only improve the recyclability and reusability performance of nano-TiO2, but also improve the photocatalytic efficiency of nano-TiO2 particles through loading to reduce the agglomeration of TiO2 particles. Meanwhile, the efficient utilization of CaCO3 mineral can be achieved by increasing its added value. Generally, TiO2-loaded mineral composite particles can be prepared by hydrolytic deposition, sol–gel and mechanical grinding methods. Sun et al. used purified diatomite as the carrier and prepared diatomite–nano-TiO2 composite photocatalyst by the hydrolysis of TiCl4 on its surface and the calcination of the product [20]. The maximum photocatalytic degradation efficiency of Rhodamine B (10 ppm) reached up to around 95% within 20 min of irradiation. Kun et al. prepared montmorillonite–nano-TiO2 composite photocatalyst by the sol–gel method with montmorillonite as the carrier [21]. The specific conversion of contaminant by the composite is three times that of TiO2. As for the loading of TiO2 on CaCO3, Tanabe [22] prepared CaCO3–TiO2 composite particles through carbonation in a TiO2 system. Sun [23] prepared CaCO3–TiO2 composite particles by the hydrophobic agglomeration method. However, these methods are complex and costly. Sun et al. prepared CaCO3–TiO2 composite pigment by liquid phase mechanical chemistry with CaCO3 as the carrier. However, this process requires a large amount of water, and subsequent processes such as filtration and drying are also required, which are complicated and consume more energy. Therefore, in this study, grinding depolymerization, dispersion and particle composition were integrated; the CaCO3–TiO2 composite photocatalyst was prepared by directly dry grinding CaCO3 and nano-TiO2 powder in a grinding machine, which has the advantages of requiring a simple process and low cost. In this study, the preparation conditions of the CaCO3–TiO2 composite photocatalyst were optimized. The morphology and structure of CaCO3–TiO2 composite particles were characterized and the photocatalytic degradation of CaCO3–TiO2 was carried out. The photocatalytic degradation of methyl orange by CaCO3–TiO2, including its reusability, was tested. Moreover, the combination mechanism between CaCO3 and TiO2 and the improvement mechanism for the photocatalytic performance of nano-TiO2 were discussed.

2. Methods

2.1. Raw Materials and Reagents

The CaCO3 raw material was calcite grinding powder produced in Jilin province, China, with a purity of 100% and whiteness of 95%. The D97 (particles with size ranging from 0 to D97 account for 97% of the weight ratio) of the used CaCO3 was 25 μm, 15 μm and 10 μm. The raw material of nano-TiO2 used in this study was the commercial degussa P25 product, which is composed of the mixed phase of anatase and rutile.

2.2. Preparation of CaCO3–TiO2 Composite Photocatalyst

According to the composite ratio, a certain amount of nano-TiO2 powder and CaCO3 powder with specific particle sizes were put into an oven and dried at 80 °C for 2 h. Then, the nano-TiO2 powder and CaCO3 powder were well mixed, and the mixed powder was added to an RK/XPM-Φ120 grinding machine. After grinding for 30 min (the diameter of grinding bowl was 120 mm, the speed was 9 r/min and the speed of the grinding rod was 220 r/min), the CaCO3–nano-TiO2 composite photocatalyst was obtained.

2.3. Photocatalytic Properties Test

The photocatalytic performance of CaCO3–TiO2 composite photocatalyst was tested with methyl orange as the target degradation product. A 300 W mercury lamp with a dominant wavelength of 254 nm was used as the light source. First, 50 mg of CaCO3–TiO2 composite photocatalyst was added to 50 mL of as-prepared methyl orange solution (the concentration was marked as C0, 10 mg/L) to obtain a suspension. In order to reduce the measurement error caused by sample adsorption, dark reaction was carried out for 1 h. After turning on the light source, the residual concentration of methyl orange was tested at specified intervals and marked as C. The photocatalytic degradation performance of the samples was characterized and evaluated by the change of C/C0. Linear regression was performed on the relationship between –Ln (C/C0) and t (lighting time), and the linear slope was used to characterize the photocatalytic degradation rate.
The measurement of the concentration of methyl orange solution after photodegradation was as follows: the solution was centrifugated, the supernatant was removed and its absorbance was measured by a UV-VIS spectrophotometer (Cary 5000, Varian company, Palo Alto, CA, USA). The concentration of methyl orange in the solution was obtained through the relationship between the absorbance and concentration.
After the photocatalytic degradation of methyl orange, the suspension was centrifuged to obtain the sediments (photocatalyst). After being dried at a low temperature, the composite photocatalyst was tested for the photocatalytic degradation performance again, which was regarded as one cycle. The above process was repeated multiple times to obtain the cycle degradation performance of the samples.

2.4. Characterization

An X-ray powder diffractometer (D/MAX2000, Rigaku Corporation, Tokyo, Japan) was used to analyze the phase composition of the CaCO3–TiO2 composite photocatalyst. A scanning electron microscope (SEM, S-3500N, HITACHI, Tokyo, Japan) was used to observe the sample morphology. The UV-VIS diffuse reflection absorption spectrum (Cary 5000, USA Varian, Palo Alto City, CA, USA) was used to characterize the light absorption properties of the samples. Fourier transform infrared spectroscopy (Spectrum 100, PerkinElmer Instruments (Shanghai) Co., Ltd., Shanghai, China) was used to test the binding properties of CaCO3 and TiO2.

3. Results and Discussion

3.1. Photocatalytic Degradation of Methyl Orange by CaCO3–TiO2

3.1.1. Effect of Particle Size of CaCO3

In order to investigate the effect of the particle size of CaCO3 on the properties of the prepared CaCO3–TiO2 composite photocatalyst, CaCO3–TiO2 composite photocatalysts were prepared with 25 μm, 15 μm and 10 μm CaCO3 as the carrier (the mass ratio of nano-TiO2 was 40%). Figure 1 shows the photocatalytic degradation of methyl orange by CaCO3–TiO2 and pure TiO2, and a blank experiment was also conducted. It can be seen that when the pure methyl orange solution was illuminated by the ultraviolet light only, the C/C0 remained almost unchanged with the increase of illumination time, indicating that there was no degradation effect on the methyl orange. By contrast, treatment with the CaCO3–TiO2 composite photocatalyst (the size of CaCO3 was 15 μm) and pure TiO2 both resulted in the effective degradation of methyl orange. After 30 min of irradiation, the C/C0 was reduced to less than 0.05, indicating that the degradation efficiency was higher than 95%; after 40 min of irradiation, the degradation efficiency reached 100%. Obviously, CaCO3–TiO2 has excellent photocatalytic performance, which is similar to that of nano-TiO2. Since the mass ratio of TiO2 to CaCO3–TiO2 was only 40%, the above results indicate that CaCO3 has an improving effect on the performance of TiO2.
Figure 1 also shows that among the three CaCO3–TiO2 samples, the composite with 15 μm CaCO3 as the carrier exhibited the highest degradation efficiency. After 20 min of irradiation, the degradation efficiency of methyl orange was 90%, and the samples with 25 μm and 10 μm CaCO3 as the carrier only reached about 82%. When the irradiation time increased to 30 min, the methyl orange degradation efficiencies of the three samples were 98%, 86% and 89%, respectively. The above results indicate that the particle size and specific surface area of CaCO3 are important, and that CaCO3 with a particle size of 15 μm properly matched with the nanoTiO2 particles to form an optimal composite relationship. Therefore, CaCO3 with a particle size of 15 μm was chosen as the carrier to be compounded with nano-TiO2.

3.1.2. Effect of Loading Amount of Nano-TiO2

As the photocatalytic active component in CaCO3–TiO2, the loading amount of nano-TiO2 should have a significant effect on the performance of the composite photocatalyst. Therefore, using 15 μm CaCO3 as the carrier, CaCO3–TiO2 composite photocatalysts with different mass ratios of nano-TiO2 were prepared, and the degradation performance of methyl orange under ultraviolet light was tested. The results are shown in Figure 2. As the mass ratio of nano-TiO2 increased from 10% to 40%, the degradation efficiency of CaCO3–TiO2 increased gradually. When the composite photocatalyst with a TiO2 mass ratio of 40% was illuminated for 20 min, the C/C0 was lower than 0.1, indicating that the degradation efficiency reached higher than 90%. And when the irradiation time increased to 40 min, the degradation efficiency reached 100%, which is equivalent to pure TiO2. When the mass ratio of nano-TiO2 increased to 50%, although the degradation efficiency was improved after 10 min of illumination, the degradation efficiency remained almost unchanged when the irradiation time was extended, with a maximum of only 90%. Obviously, the mass ratio of nano-TiO2 in CaCO3–TiO2 should be set as 40%.
Since CaCO3–TiO2 with only a small mass ratio of nano-TiO2 reaches the photocatalytic degradation effect similar to that of pure nano-TiO2, the use amount of TiO2 can be significantly reduced and thus cost can be saved in practical applications. This is a result of the significant improvement in the photocatalytic efficiency of TiO2 by its combination with CaCO3. It is speculated that this improvement results from the improvement of the dispersibility of nano-TiO2 and the formation of a combination interface between TiO2 and CaCO3.

3.1.3. Photocatalytic Degradation of Different Concentrations of Methyl Orange Solution by CaCO3–TiO2

The investigation of the degradation performance of CaCO3–TiO2 in different concentrations of methyl orange solution can not only further verify the photocatalytic degradation ability of CaCO3–TiO2, but it can also determine its reasonable scope of application. Figure 3 shows the photocatalytic degradation effect of CaCO3–TiO2 for different concentrations of methyl orange solution. It can be seen that the degradation efficiency of the CaCO3–TiO2 composite photocatalyst on each concentration of methyl orange solution increased with the increase of the illumination time, and finally reached a higher degradation effect. Comparing the degradation effect of different concentrations, it was found that the degradation efficiency of methyl orange solution with concentrations of 10 and 20 ppm by CaCO3–TiO2 was higher than that of concentrations of 50 and 80 ppm. Among them, the methyl orange solution with concentrations of 10 and 20 ppm could be completely degraded (C/C0 = 0) by CaCO3–TiO2 at illumination times of 60 and 80 min, respectively, and the degradation rate was faster. In contrast, the degradation rates of 50 and 80 ppm methyl orange solution by CaCO3–TiO2 were reduced, and the degradation efficiencies were 90% and 80%, respectively (C/C0 was 0.1 and 0.2, respectively) at illumination times of 60 and 90 min. These results show that the CaCO3–TiO2 composite photocatalyst can effectively degrade methyl orange solution with concentrations of 10–80 ppm. The reason why the degradation rate decreases with the increasing concentration of methyl orange solution is that high concentrations of pollutants require longer reaction times.

3.1.4. Recycling Performance of CaCO3–TiO2 Composite Photocatalyst

The recycling performance of the catalyst is an important factor reflecting its stability and practical application performance. Therefore, we investigated the recycling performance of the optimum CaCO3–TiO2 composite photocatalyst (the particle size of CaCO3 was 15 μm, the mass ratio of TiO2 was 40%). The results are shown in Figure 4. The results show that the photocatalytic degradation performance of CaCO3–TiO2 after five cycles is similar to that of the first use, indicating that recycling does not reduce its performance. This indicates that the CaCO3–TiO2 composite photocatalyst prepared in this study has excellent recycling performance and can be recycled and reused several times, thereby reducing the cost. Obviously, this is due to the loading of TiO2 on the surface of CaCO3, which results in the nano-TiO2 being firmly fixed on the surface of CaCO3. Therefore, the loss of TiO2 is prevented, and CaCO3–TiO2 can easily be separated from water due to its large particle size. However, single nano-TiO2 particles have a small particle size, so they are difficult to precipitate and easily lost when separated from their carrier. Therefore, it is believed that the CaCO3–TiO2 composite photocatalyst has practical application prospects for sewage treatment.

3.2. Optical Performance Test

Figure 5 shows the UV-VIS diffuse reflectance absorption spectra (DRS) of the CaCO3–TiO2 composite photocatalyst as well as the raw materials of CaCO3 and TiO2. It can be seen that CaCO3–TiO2 and TiO2 exhibit similar light absorption characteristics. They show no light absorption in the visible light region (wavelengths of 400–800 nm) and show strong absorption in ultraviolet light (wavelengths of 200–400 nm), which is consistent with the bandgap characteristics of the TiO2 semiconductor (3.2 eV of bandgap width). By contrast, CaCO3 shows almost no absorption of both visible light and ultraviolet light. The results show that CaCO3–TiO2 has similar properties to TiO2, consistent with the observed strong photocatalytic degradation performance of CaCO3–TiO2 and TiO2. Meanwhile, it is inferred that a composite structure of CaCO3 with surface-loaded nano-TiO2 is formed in the creation of CaCO3–TiO2.

3.3. Morphology and Structure of CaCO3–TiO2 Composite Photocatalyst

3.3.1. SEM Analysis

Figure 6 shows SEM images of CaCO3, nano-TiO2 raw materials, CaCO3–TiO2 composite photocatalyst with different mass ratios of TiO2 and the physical mixtures of CaCO3 and TiO2 (40% TiO2 mass ratio). Figure 6a shows that the CaCO3 particles have a bulk shape with a particle size of about 1–3 μm, as well as smooth surfaces without coatings. Figure 6b,c shows that the nano-TiO2 particles are fine, but their agglomeration is severe, with an aggregate size of up to 2 μm. Figure 6d–g show that when CaCO3 is combined with nano-TiO2, the surface of CaCO3 particles is uniformly covered by fine particles, and the surface becomes rough. Figure 6i,j are the mapping results of Ca and Ti elements corresponding to Figure 6f; it can be seen that Ca distributes in the particle contour range of CaCO3, reflecting the characteristic of CaCO3. The distribution of Ti in the scanning area is uniform, which corresponds to the location of the particles. This proves that the TiO2 is distributed on the surface of CaCO3 uniformly. Obviously, the CaCO3–TiO2 composite photocatalyst is composed of composite particles characterized by nano-TiO2 loaded on the CaCO3 surface. Therefore, it is inferred that CaCO3–TiO2 should exhibit the properties of nano-TiO2 (photocatalytic properties). In addition, it was found that the dispersion of TiO2 loaded on the surface of CaCO3 was significantly higher than that of TiO2 alone. The particle unit size is generally reduced to less than 0.2 μm, which allows the active sites of TiO2 to be more exposed, thereby increasing its photocatalytic performance. This is undoubtedly the intrinsic mechanism enabling CaCO3–TiO2 to exhibit a photocatalytic degradation performance comparable to that of pure TiO2 when the mass ratio of TiO2 is only 40%.
Figure 6d–f shows that as the mass ratio of TiO2 increased from 20% to 50%, the surface of the CaCO3 was coated by nano-TiO2 particles more and more uniformly and completely. Especially when the mass ratio of TiO2 was 40%, the surface of CaCO3 was almost completely covered by TiO2 particles. However, in the physical mixture of CaCO3 and TiO2 with the same composite proportion (Figure 6h), the loading of nano-TiO2 was very poor. Only a small part of the surface of CaCO3 particles contained loadings, and most of them were bare. The above results indicate that the co-grinding of CaCO3 and TiO2 greatly promotes the loading function, thereby improving the performance of CaCO3–TiO2.

3.3.2. XRD Analysis

Figure 7 displays the XRD spectra of CaCO3, nano-TiO2 raw material, CaCO3–TiO2 composite photocatalyst and the physical mixture of CaCO3 and TiO2. The results reveal that only the characteristic diffraction peaks of calcite appeared in the XRD spectrum of CaCO3, indicating its high purity. The characteristic diffraction peaks of rutile and anatase appeared in the spectrum of nano-TiO2, indicating that it consisted of rutile and anatase mixed phase, consistent with the phase composition of P25. In the spectrum of the CaCO3–TiO2 composite photocatalyst, only calcite, rutile and anatase crystal phases were observed, indicating that the compounding of CaCO3 and TiO2 does not produce a new phase. It was concluded that the combination of CaCO3 and TiO2 occurs in their interface region, and the binding properties should be chemical or physical action within the interface area. The phase composition of the physical mixture of CaCO3 and TiO2 was found to be the same as that of CaCO3–TiO2 composite photocatalyst, but the intensity of the characteristic peaks of CaCO3 in CaCO3–TiO2 composite photocatalyst was weaker than that in the physical mixture. This may be due to the reduction of CaCO3 exposure in CaCO3–TiO2 due to the coating of the surface with titanium dioxide, while most of the surface of CaCO3 in the physical mixture was still bare [23]. The above results also indicate the successful compounding of CaCO3 and TiO2 in CaCO3–TiO2.

3.3.3. Infrared Spectrum Analysis

The infrared spectra of CaCO3, nano-TiO2 raw material and CaCO3–TiO2 composite photocatalyst were measured, as shown in Figure 8. In the infrared spectrum of nano-TiO2, the wide absorption band in the range of 3250–3700 cm−1 and the absorption peak at 1636 cm−1 respectively represent the stretching vibration and bending vibration of the O–H bond in the water molecules and hydroxyl groups on the surface of TiO2. The absorption band in the range of about 3300–3500 cm−1 in the CaCO3 spectrum represents the characteristics of adsorption water and hydroxyl groups on its surface. In the infrared spectrum of the CaCO3–TiO2 composite photocatalyst, the absorption peak at 1636 cm−1 of TiO2 disappeared, and the absorption peak of CaCO3 ranging from 3300 to 3500 cm−1 was weakened. It has been suggested that CaCO3 and TiO2 form a firm chemical combination through their surface hydroxyl groups [24].
Based on the above study, the mechanism of CaCO3 as a carrier to improve the photocatalytic performance and utilization efficiency of nano-TiO2 can be summarized as follows: (1) The dispersibility of TiO2 is improved by uniformly loading TiO2 on the surface of CaCO3, thereby causing an increase in active sites and light absorption areas. (2) A chemical bond is formed between TiO2 and CaCO3 at their interface, which provides a new transport channel for the photo electrons [25,26], and further improves the separation efficiency of electrons and holes. (3) In CaCO3–TiO2, nano-TiO2 and CaCO3 are firmly combined due to their surface chemical bonding, which prevents the loss of TiO2 when degrading pollutants in water [9,27]. Therefore, this composite photocatalyst can be easily recovered and recycled from water by sedimentation, thereby reducing the use amount of TiO2 as well as the cost. Figure 9 is a schematic diagram reflecting the above principles.

4. Conclusions

(1)
The CaCO3–TiO2 composite photocatalyst was prepared by the co-grinding of CaCO3 and TiO2. The optimal CaCO3–TiO2 composite photocatalyst, in which the particles size of CaCO3 is 15 μm and the mass ratio of TiO2 is 40%, shows excellent photocatalytic degradation performance towards methyl orange and good recovery performance. The degradation efficiency of optimal CaCO3–TiO2 composite photocatalyst was found to be 90% and 100% after 20 and 40 min of ultraviolet light illumination, respectively. The degradation effect is comparable to pure TiO2. Moreover, its degradation effect on methyl orange is not significantly reduced after five cycles.
(2)
CaCO3–TiO2 composite photocatalyst is characterized by CaCO3 loaded by nano-TiO2 uniformly and completely. The dispersibility of the loaded TiO2 is significantly enhanced compared to pure TiO2, and a strong chemical bond is formed between CaCO3 and TiO2 particle interfaces. These are important mechanisms for improving the photocatalytic efficiency of nano-TiO2 and reducing the its amount in CaCO3–TiO2.

Author Contributions

Conceptualization, H.D. and J.W.; methodology, J.W.; software, S.S.; validation, J.W., S.S. and L.P.; formal analysis, H.D.; investigation, L.P.; resources, H.D.; data curation, J.W.; writing—original draft preparation, L.P. and Z.X.; writing—review and editing, J.W.; visualization, S.S.; supervision, W.L.; project administration, H.D.; funding acquisition, H.D.

Funding

This research was funded by Graduation support project (entrepreneurship) of Beijing university in 2018, sp-2018-300.

Acknowledgments

This work was supported by the Graduation support project (entrepreneurship) of Beijing university in 2018 (sp-2018-300).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  2. Ibhadon, A.O.; Fitzpatrick, P. Heterogeneous photocatalysis: Recent advances and applications. Catalysts 2013, 3, 189–218. [Google Scholar] [CrossRef]
  3. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photoch. C 2012, 13, 169–189. [Google Scholar] [CrossRef]
  4. Fagan, R.; Mccormack, D.E.; Dionysiou, D.D.; Pillai, S.C. A review of solar and visible light active TiO2 photocatalysis for treating bacteria, cyanotoxins and contaminants of emerging concern. Mat. Sci. Semicon. Proc. 2016, 42, 2–14. [Google Scholar] [CrossRef]
  5. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  6. Sun, B.; Reddy, E.P.; Smirniotis, P.G. Effect of the Cr6+ concentration in Cr-incorporated TiO2-loaded MCM-41 catalysts for visible light photocatalysis. Appl. Catal. B Environ. 2005, 57, 139–149. [Google Scholar] [CrossRef]
  7. Wei, Z.; Liang, F.; Liu, Y.; Luo, W.; Wang, J.; Yao, W.; Zhu, Y. Photoelectrocatalytic degradation of phenol-containing wastewater by TiO2/g-C3N4 hybrid heterostructure thin film. Appl. Catal. B Environ. 2017, 201, 600–606. [Google Scholar] [CrossRef]
  8. Su, Y.F.; Wang, G.B.; Kuo, D.T.F.; Chang, M.L.; Shih, Y.H. Photoelectrocatalytic degradation of the antibiotic sulfamethoxazole using TiO2/Ti photoanode. Appl. Catal. B Environ. 2016, 186, 184–192. [Google Scholar] [CrossRef]
  9. Zhang, G.; Sun, Z.; Duan, Y.; Ma, R.; Zheng, S. Synthesis of nano-TiO2/diatomite composite and its photocatalytic degradation of gaseous formaldehyde. Appl. Sur. Sci. 2017, 412, 105–112. [Google Scholar] [CrossRef]
  10. Wang, B.; Zhang, G.; Sun, Z.; Zheng, S. Synthesis of natural porous minerals supported TiO2 nanoparticles and their photocatalytic performance towards Rhodamine B degradation. Powder Technol. 2014, 262, 1–8. [Google Scholar] [CrossRef]
  11. Zelekew, O.A.; Kuo, D.H.; Yassin, J.M.; Ahmed, K.E.; Abdullah, H. Synthesis of efficient silica supported TiO2/Ag2O heterostructured catalyst with enhanced photocatalytic performance. Appl. Sur. Sci. 2017, 410, 454–463. [Google Scholar] [CrossRef]
  12. Kibanova, D.; Trejo, M.; Destaillats, H.; Cervini-Silva, J. Synthesis of hectorite–TiO2 and kaolinite–TiO2 nanocomposites with photocatalytic activity for the degradation of model air pollutants. Appl. Clay Sci. 2009, 42, 563–568. [Google Scholar] [CrossRef]
  13. Reddy, E.P.; Sun, B.; Smirniotis, P.G. Transition metal modified TiO2-loaded MCM-41 catalysts for visible-and UV-Light driven photodegradation of aqueous organic pollutants. J. Phys. Chem. B 2004, 108, 17198–17205. [Google Scholar] [CrossRef]
  14. Xia, Y.; Li, F.; Jiang, Y.; Xia, M.; Xue, B.; Li, Y. Interface actions between TiO2 and porous diatomite on the structure and photocatalytic activity of TiO2-diatomite. Appl. Sur. Sci. 2014, 303, 290–296. [Google Scholar] [CrossRef]
  15. Anandan, S.; Yoon, M. Photocatalytic activities of the nano-sized TiO2-supported Y-zeolites. J. Photochem. Photobiol. C 2003, 4, 5–18. [Google Scholar] [CrossRef]
  16. Mamulová, K.K.; Tokarský, J.; Kovář, P.; Vojtěšková, S.; Kovářová, A.; Smetana, B.; Kukutschová, J.; Čapková, P.; Matějka, V. Preparation and characterization of photoactive composite kaolinite/TiO2. J. Hazard. Mater. 2011, 188, 212–220. [Google Scholar] [CrossRef]
  17. Chen, D.; Zhu, Q.; Zhou, F.; Deng, X.; Li, F. Synthesis and photocatalytic performances of the TiO2 pillared montmorillonite. J. Hazard. Mater. 2012, 235, 186–193. [Google Scholar] [CrossRef]
  18. Matschei, T.; Lothenbach, B.; Glasser, F.P. The role of calcium carbonate in cement hydration. Cem. Concr. Res. 2007, 37, 551–558. [Google Scholar] [CrossRef]
  19. Poh, C.L.; Mariatti, M.; Fauzi, M.A.; Ng, C.H.; Chee, C.K.; Chuah, T.P. Tensile, dielectric, and thermal properties of epoxy composites filled with silica, mica, and calcium carbonate. J. Mater. Sci. Mater. Electroc. 2014, 25, 2111–2119. [Google Scholar] [CrossRef]
  20. Sun, Z.; Hu, Z.; Yan, Y.; Zheng, S. Effect of preparation conditions on the characteristics and photocatalytic activity of TiO2/purified diatomite composite photocatalysts. Appl. Surf. Sci. 2014, 314, 251–259. [Google Scholar] [CrossRef]
  21. Kun, R.; Mogyorósi, K.; Dékány, I. Synthesis and structural and photocatalytic properties of TiO2/montmorillonite nanocomposites. Appl. Clay Sci. 2006, 32, 99–110. [Google Scholar] [CrossRef]
  22. Tanabe, K.; Anabe, K.; Minitsuhashi, K.; Yoshida, T. Titanium Dioxide–Calcium Carbonate Composite Particles. U.S. Patent 6,991,677, 31 January 2006. [Google Scholar]
  23. Sun, S.; Ding, H.; Hou, X.; Chen, D.; Yu, S.; Zhou, H.; Chen, Y. Effects of organic modifiers on the properties of TiO2-coated CaCO3 composite pigments prepared by the hydrophobic aggregation of particles. Appl. Surf. Sci. 2018, 456, 923–931. [Google Scholar] [CrossRef]
  24. Sun, S.; Ding, H.; Hou, X. Preparation of CaCO3-TiO2 composite particles and their pigment properties. Materials 2018, 11, 1131. [Google Scholar] [CrossRef] [PubMed]
  25. Li, C.; Sun, Z.; Zhang, W.; Yu, C.; Zheng, S. Highly efficient g-C3N4/TiO2/kaolinite composite with novel three-dimensional structure and enhanced visible light responding ability towards ciprofloxacin and S. aureus. Appl. Catal. B Environ. 2017, 220, 272–282. [Google Scholar] [CrossRef]
  26. Lin, S.; Zhang, Y.; You, Y.; Zeng, C.; Xiao, X.; Ma, T.; Huang, H. Bifunctional Hydrogen Production and Storage on 0D–1D Heterojunction of Cd0.5Zn0.5S@Halloysites. Adv. Funct. Mater. 2019, 1903825. [Google Scholar] [CrossRef]
  27. Li, G.; Park, S.; Rittmann, B.E. Developing an efficient TiO2-coated biofilm carrier for intimate coupling of photocatalysis and biodegradation. Water. Res. 2012, 46, 6489–6496. [Google Scholar] [CrossRef]
Figure 1. Effect of the particle size of CaCO3 on the performance of the composite photocatalyst.
Figure 1. Effect of the particle size of CaCO3 on the performance of the composite photocatalyst.
Materials 12 03369 g001
Figure 2. Effect of mass ratio of TiO2 on the properties of the composite photocatalyst.
Figure 2. Effect of mass ratio of TiO2 on the properties of the composite photocatalyst.
Materials 12 03369 g002
Figure 3. Effect of the concentration of methyl orange solution on the degradation efficiency (a) and degradation rate (b) of CaCO3–TiO2 composite photocatalyst.
Figure 3. Effect of the concentration of methyl orange solution on the degradation efficiency (a) and degradation rate (b) of CaCO3–TiO2 composite photocatalyst.
Materials 12 03369 g003
Figure 4. Cyclic curve of the degradation of methyl orange by CaCO3–TiO2 composite photocatalyst.
Figure 4. Cyclic curve of the degradation of methyl orange by CaCO3–TiO2 composite photocatalyst.
Materials 12 03369 g004
Figure 5. UV-VIS absorption spectra of CaCO3, TiO2 and CaCO3–TiO2.
Figure 5. UV-VIS absorption spectra of CaCO3, TiO2 and CaCO3–TiO2.
Materials 12 03369 g005
Figure 6. SEM images of CaCO3 (a), TiO2 (b,c), CaCO3–TiO2 composite photocatalyst with TiO2 mass ratios of 20% (d), 40% (e), 50% (f,g). The mixture of CaCO3 and TiO2 (h). The mapping of Ca (i) and Ti (j) corresponding to (f).
Figure 6. SEM images of CaCO3 (a), TiO2 (b,c), CaCO3–TiO2 composite photocatalyst with TiO2 mass ratios of 20% (d), 40% (e), 50% (f,g). The mixture of CaCO3 and TiO2 (h). The mapping of Ca (i) and Ti (j) corresponding to (f).
Materials 12 03369 g006
Figure 7. X-ray diffraction (XRD) spectra of CaCO3, TiO2, CaCO3–TiO2 composite and the physical mixture of CaCO3 and TiO2.
Figure 7. X-ray diffraction (XRD) spectra of CaCO3, TiO2, CaCO3–TiO2 composite and the physical mixture of CaCO3 and TiO2.
Materials 12 03369 g007
Figure 8. FT-IR spectra of CaCO3, TiO2 and CaCO3–TiO2 composite.
Figure 8. FT-IR spectra of CaCO3, TiO2 and CaCO3–TiO2 composite.
Materials 12 03369 g008
Figure 9. Improvement mechanism of photocatalytic performance.
Figure 9. Improvement mechanism of photocatalytic performance.
Materials 12 03369 g009

Share and Cite

MDPI and ACS Style

Wang, J.; Sun, S.; Pan, L.; Xu, Z.; Ding, H.; Li, W. Preparation and Properties of CaCO3-Supported Nano-TiO2 Composite with Improved Photocatalytic Performance. Materials 2019, 12, 3369. https://doi.org/10.3390/ma12203369

AMA Style

Wang J, Sun S, Pan L, Xu Z, Ding H, Li W. Preparation and Properties of CaCO3-Supported Nano-TiO2 Composite with Improved Photocatalytic Performance. Materials. 2019; 12(20):3369. https://doi.org/10.3390/ma12203369

Chicago/Turabian Style

Wang, Jie, Sijia Sun, Lei Pan, Zhuoqun Xu, Hao Ding, and Wei Li. 2019. "Preparation and Properties of CaCO3-Supported Nano-TiO2 Composite with Improved Photocatalytic Performance" Materials 12, no. 20: 3369. https://doi.org/10.3390/ma12203369

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop