Effect of heating time on structural, morphology, optical, and photocatalytic properties of g-C3N4 nanosheets
Hung Nguyen Manh1, 4, Mai Oanh Le Thi1, 2, †, Hang Lam Thi1, 3, Chung Pham Do2, Duyen Pham Thi1, 5, Thang Dao Viet1, 4, Van Minh Nguyen1, 2
Center for Nano Science and Technology, Hanoi National University of Education, 136 Xuan Thuy Road, Cau Giay District, Hanoi, 100000, Vietnam
Department of Physics, Hanoi National University of Education, 136 Xuan Thuy Road, Cau Giay District, Hanoi, 100000, Vietnam
Faculty of Basic Sciences, Hanoi University of Natural Resources and Environment, 41A Phu Dien Road, North Tu Liem, Hanoi, 100000, Vietnam
Hanoi University of Mining and Geology, Duc Thang ward, North Tu Liem District, Hanoi, 100000, Vietnam
Military science Academy, 322 Le Trong Tan street, Dinh Cong, Hoang Mai, Hanoi, 100000, Vietnam

 

† Corresponding author. E-mail: lemaioanh@gmail.com

Project supported by the scientific and technological project at the level of Ministry of Education, Vietnam (Grant No. B2018-SPH-06-CTrVL).

Abstract

Effect of heating time on the structural, morphology, optical, and photocatalytic properties of graphitic carbon nitride (g-C3N4) nanosheets prepared at 550 °C in Ar atmosphere is studied. The investigations are carried out by using x-ray diffraction (XRD), scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FTIR), UV–vis absorption, and photoluminescence (PL). At a heating temperature of 550 °C, g-C3N4 nanocrystals are formed after 0.5 h and become more orderly as the heating time increases. The surface area of the g-C3N4 nanosheets significantly increases as the preparation time increases. The g-C3N4 prepared in 2.5 h shows the highest photocatalytic performance, decomposing completely 10 ppm RhB solution under xenon lamp irradiation for 2.0 h.

PACS: ;71.20.NR;;71.20.RV;
1. Introduction

Facing the exhaustion of fossil fuel resources and serious environmental pollution due to industrial and domestic wastes, scientists are actively looking for efficient photocatalysts which can convert solar energy into chemical energy and be used in the photodegradation of environmental pollutants, hydrogen production via water splitting, air treatment, and CO2 conversion to hydrocarbon fuel.[15] In the last few years, graphitic carbon nitride (g-C3N4), as a new material, has come into being in this area, and it is a non-metallic semiconductor with a band gap energy of about 2.7 eV (λ ∼ 460 nm), durable, non-toxic, highly recyclable and exhibits superior photocatalytic ability.[2,610] Therefore, g-C3N4 has become an attractive research object in the field of photocatalyst.

To obtain g-C3N4, it is necessary to use carbon and nitrogen-rich precursors such as urea,[11,12] thiourea,[11] melamine,[13,14] cyanamide,[15] dicyandiamide,[16] etc. Common methods for preparing g-C3N4 include high-pressure pyrolysis,[17] thermal polymerization,[14] hydrothermal,[18,19] etc. One of the C and N-rich precursors, not expensive and easy to find, is urea. Our previous study showed that with a heating time of 2 h, g-C3N4 nanosheets prepared by thermal polymerization in Ar atmosphere at 550 °C exhibited the highest photocatalytic efficiency.[20] In the present work, the heating temperature is fixed at 550 °C while the heating time is changed from 0.5 to 2.5 h to study the effect of the heating time on the formation of crystal structure, optical property, and photocatalytic performance of g-C3N4 nanosheets.

2. Experiment
2.1. Preparation of g-C3N4

Thermal polymerization in inert Ar atmosphere was adopted to synthesize g-C3N4. Firstly, 5 g urea (≥ 97.0%, Sigma-Aldrich) was mixed with ethanol solution, ground for 1 h, and transferred into an alumina crucible. Thermal polymerization process took place at the heating temperature of 550 °C for different time from 0.5 h to 2.5 h in Ar atmosphere by using a tubular furnace. The samples are denoted as CN-x (x = 0.5, 1.0, 1.5, 2.0, and 2.5) according to their preparation time.

2.2. Photocatalytic degradation experiment

The 10 ppm Rhodamine B (RhB) solution was used as a reagent to determine the photocatalytic activity of the as-synthesized g-C3N4. Initial RhB solution was prepared at a concentration of 20 ppm. As the first step, 0.06 g g-C3N4 nanosheets was dispersed in 30 ml distilled water under ultrasonic vibration for 1 h. In the second step, 30 ml g-C3N4-containing solution was mixed with 30 ml 20-ppm RhB solution to obtain 60 ml 10-ppm RhB solution containing g-C3N4.

Firstly, the RhB adsorption experiment was carried out to assess the adsorption ability of g-C3N4. The 10 ppm RhB solution containing g-C3N4 was stirred in a dark chamber for 30 min to reach an adsorption–desorption equilibrium. The solution was then illuminated by a 300 mW xenon lamp. The xenon lamp was placed at a distance of 6 cm from the surface of the RhB solution. After every 15 min of illumination, an amount of 3 ml of the suspension was taken and centrifuged to remove g-C3N4 for UV–vis absorption measurement. The intensity of the 552 nm absorption peak was used to determine the remaining RhB concentration in the solution from which the C/C0 ratio was calculated.

2.3. Characterizations

The D8 Advance diffractometer (Bruker) with Cu-Kα radiation was used to record the x-ray diffraction (XRD) patterns of the as-synthesized g-C3N4 nanosheets. The morphologies of the as-synthesized g-C3N4 nanosheets were observed by using a scanning electron microscopy (SEM) (S-4800 NIHE microscope, Hitachi). The Brunauer–Emmett–Teller (BET) surface area was measured by a Micromeritics 3Flex. Ultraviolet–visible (UV–vis) absorption spectra were measured by a Jasco V670 UV–vis spectrophotometer through the diffuse reflectance spectroscopy (DRS) technique for the g-C3N4 powder. Fourier transform infrared spectra (FTIR) were measured by an IR Prestige-21 FTIR/NIR spectrometer (Shimadzu). Photoluminescence (PL) spectra were also measured by a fluorescence spectrophotometer (Nanolog iHR 320, Horiba) at the excitation wavelength of 350 nm.

3. Results and discussion

The XRD patterns of the g-C3N4 samples prepared at the heating temperature of 550 °C for different time are shown in Fig. 1(a). The samples all exhibit three diffraction peaks located at 12.7°,24.8°, and 27.5° which are corresponding to the (001), (101), and (002) diffraction planes, respectively. The peak intensity increases as the heating time increases. A comparison between these peaks and the (001), (101), and (002) peaks of standard g-C3N4 materials demonstrates that the synthesized g-C3N4 is crystallized in hexagonal phase (JCPDS file No. 87-1526). The strongest peak at 27.5° corresponds to the interlayer stacking of the conjugated aromatic system[21] with an interlayer distance of 3.42 Å. It is obvious that as the heating times increase, the (002) peak shifts slightly towards the larger diffraction angle, indicating the narrowing of the interlayer distance and the decrease of crystal parameter c. The calculations show that parameter c decreases from 6.85 Å for CN-0.5 sample to 6.77 Å for CN-2.5 sample (Table 1). The lattice parameter a can be calculated via peak (101) located at 24.8°. The calculations show that parameter a decreases from 4.59 Å to 4.55 Å when the heating time increases from 0.5 h to 2.5 h (Table 1). Parameters a and c varying with the heating time are plotted in Fig. 1(b), which shows that th eg-C3N4 crystal is packed more tightly as the processing time increases, similar to what happens as the heating temperature increases.[20] In addition, figure 1(a) shows that the diffraction intensity increases with the peak width narrowing when the heating time increases (Table 1), indicating that the polymerization of g-C3N4 takes place more strongly and crystallinity of g-C3N4 is higher as the heating time increases. The XRD peak position and peak width change noticeably as the heating time increases from 1.0 h to 1.5 h, which shows that a fairly ordered lattice can be achieved after heating urea at 550 °C for 1.5 h.

Fig. 1. (a) XRD patterns of g-C3N4 samples heated at 550 °C for different heating time and (b) lattice parameters varying with heating time.
Table 1.

Peak positions, heights, and full widths at half maximum (FWHM) of (002) XRD peaks, and lattice parameters a and c of as-prepared samples.

.

Figure 2 shows the infrared absorption spectra of the as-synthesized g-C3N4 samples with different heating time. According to the previous studies, the breathing modes of CN heterocycles of triazine units are responsible for the absorption peak at 810 nm while the N–H group substituted into aromatic primary amine usually shows the stretching absorption in the region of 3000–3600 nm. The stretching mode of the C = N bonds in aromatic rings is responsible for the absorption peak at 1640 cm−1 while the stretching modes of the C–N bonds cause some absorption bands in the range of 1200–1500 cm−1. The FTIR spectra of the g-C3N4 samples change slightly as the preparation time increases as indicated by the leftward shifting of the peaks 1242 cm−1 and 1413 cm−1 and the rightward shifting of the peaks 1321 cm−1 and 1457 cm−1 in Fig. 2(b). In the absorption region of 3000–3600 cm−1, some stretching frequencies are observed at 3084 cm−1, 3162 cm−1, 3269 cm−1, and 3413 cm−1, among which the 3413 cm−1 and 3162 cm−1 bands can be attributed to asymmetric stretching and symmetric stretching of the NH2 group while the 3084 cm−1 and 3269 cm−1 bands are due to the O–H stretching vibrations.[22] It is obviously seen that the 3413 cm−1 band shifts gradually toward higher frequency with heating time increasing, consistent with the XRD analysis in which the crystal becomes more ordered and the crystal parameters decrease. In addition, the intensity ratio of 3413 cm−1 to 3162 cm−1 first increases, and then decreases. This also strengthens the change in crystal structure of g-C3N4 as heating time increases.

Fig. 2. (a) FTIR spectra of g-C3N4 samples prepared at 550 °C with different preparation time and (b) FTIR peak positions varying with wavelength.

The SEM images of g-C3N4 synthesized with different heating time are shown in Fig. 3. The g-C3N4 sample obtained at 550 °C for 0.5 h in Fig. 3(a) shows heterogeneous particle morphology. Some particles are granular while the others are in the form of nanosheets with a thickness of around 10 nm. When the heating time increases to 1.0 h, the granular particles no longer exist, leaving nanosheets with curved edges. For 1.0 h, the sheets are quite large with very few torn holes on the surface. However, when the heating time increases to 2.0 h and 2.5 h, the samples are still in the form of nanosheets, however, the surfaces of the sheets are punctured or torn into smaller sheets. The g-C3N4 obtained at 550 °C for 2.5 h consists of nanosheets that are small in size and more porous than the other samples, which is beneficial for photocatalytic performance due to its large specific surface area.

Fig. 3. FE-SEM images of g-C3N4 nanosheets synthesized at 550 °C with different preparation time: (a) 0.5 h, (b) 1.0 h, (c) 2.0 h, and (d) 2.5 h.

The BET technique is used to measure the specific surface area and pore size distributions of the as-prepared g-C3N4 samples with heating time of 1.0 h, 2.0 h, and 2.5 h as shown in Fig. 4. All samples show type IV isotherms with H3 hysteresis loops, indicating a mesoporous structure that forms slit-shaped pores with a non-uniform size and/or shape. The calculated surface area, average pore volume, and average pore size are listed in Table 2. The surface area of the samples increases from 60 m2/g to 71 m2/g and 88 m2/g for CN-1.0, CN-2.0, and CN-2.5, respectively, which are larger than the specific surface area of pure g-C3N4 reported previously.[23,24] This difference is due to the nanosheets being punctured and torn as the heating time increases as seen in the SEM images. All samples exhibit a wide pore size distribution (Fig. 4(b)) and a large average pore width in a range of 10–30 nm.

Fig. 4. (a) N2 adsorption–desorption isotherms of g-C3N4 heated for 1.0 h, 2.0 h, and 2.5 h and (b) plots of BJH pore volume versus pore diameter.
Table 2.

Surface areas, pore volumes, and pore sizes of the as-prepared samples.

.

UV–vis diffuse reflectance spectra (DRS) of g-C3N4 obtained with different preparation time in Fig. 5(a) indicate an absorption edge at around 450 nm for all samples. It is obvious that the absorption edge first shifts slightly to the right as the heating time increases from 0.5 h to 1.0 h and then inversely shifts to the left with the increasing preparation time. For indirect band gap semiconductor such as g-C3N4, ΔEg can be determined by using Tauc plot of (α )1/2 versus photon energy () as shown in Fig. 5(b). The band gap is 2.73 eV and 2.66 eV for CN-1.0 and CN-2.0, respectively, revealing that the visible light can be used for exciting photocatalysis.

Fig. 5. (a) Absorption spectra of g-C3N4 nanosheets obtained at 550 °C with different preparation time and (b) Wood–Tauc method determining band gap energy of indirect semiconductor.

Figure 6 shows the photoluminescence spectra PL (with excitation wavelength of 350 nm) of the as-synthesized g-C3N4 samples with different processing time, indicating a broad blue–green emission band for all samples. The PL band concentrates at 450 nm, in accordance with the UV–vis absorption edge. This PL band can be decomposed into four component peaks as shown in the inset of Fig. 6(a) (for CN-1.5 sample), including P1 (430.4 nm, 2.88 eV), P2 (453.6 nm, 2.73 eV), P3 (491.6 nm, 2.52 eV), and P4 (545.8 nm, 2.27 eV) which accord well with previous reports.[20,25] The PL peaks shift slightly towards the longer wavelength as the processing time increases (see Fig. 6(b)), consistent with the decrease of the energy bandgap obtained in the absorption spectra. In addition, the PL intensity first gradually increases to the highest value at CN-2.0 and then decreases.

Fig. 6. (a) Photoluminescence spectra of g-C3N4 nanosheets obtained at 550 °C with different preparation time, with inset showing the normalized PL spectra and dot lines being Gaussian fitting for CN-2.0 sample, and (b) PL peak positions varying with heating time.

The photocatalytic activity of the g-C3N4 sample is assessed through the degradation of Rhodamine B (RhB) solution under the irradiation of xenon lamp in which the UV–vis absorption technique was used to indirectly measure the concentration of remaining RhB. Figure 7 shows the C/C0 ratio as a function of time, indicating that the RhB degradation highly depends on the heating time. Samples CN-0.5 and CN-1.0 exhibit low photocatalytic efficiency, only about 60% of RhB is degraded after 2 h of exposing. The photocatalytic activity increases sharply as the preparing time increases to 2 h. The CN-2.0 and CN-2.5 degrade almost all RhB after 2 h of exposing. This improvement of photocatalytic performance can be explained by the mesoporous structure of the as-prepared g-C3N4 with long heating time. Although CN-2.0 has lower specific surface area than CN-2.5, its pore volume and pore size are both larger, so their photocatalytic performances are almost the same. In addition, the narrower energy band gap of the sample with long processing time also contributes to the improvement of the photocatalytic activity.

Fig. 7. Adsorption and photocatalytic activities of the g-C3N4 prepared at 550 °C with different preparation time in decomposition of RhB solution under xenon lamp irradiation.
4. Conclusion

The structural and physical properties of g-C3N4 nanosheets are greatly influenced by the heating time. The crystal lattice is formed more orderly and packed more densely as the heating time increases. The increase in fabrication time also causes the g-C3N4 nanosheets to be slightly torn, making them more porous and the specific surface area increase. The energy band gap changes negligibly as the processing time increases. However, the photoluminescence peaks shift slightly toward longer wavelength. The changes in structure, morphology, and physical properties determine the change in the photocatalytic performance of the g-C3N4 nanosheets. The g-C3N4 nanosheets obtained in long heating time, greater than 2.0 h, possess high photocatalytic activities: the RhB decomposed after 2.0 h of xenon lamp irradiating is accounted for 95%.

Reference
[1] Gopal N O Lo H H Ke T F et al. 2012 J. Phys. Chem. 116 16191
[2] Kumar S Karthikeyan S Lee A 2018 Catalysts 8 74
[3] Bi Y Ehsan M F Huang Y Jin J He T 2015 J. CO2 Utilization 12 43
[4] Zhai X Deng H Zhou W Yang P Chu J Zheng Z 2015 Mater. Lett. 161 423
[5] Bai X Wang L Zhu Y 2012 ACS Catal. 2 2769
[6] Ye S Wang R Wu M Z Yuan Y P 2015 Appl. Surf. Sci. 358 15
[7] Wen J Xie J Chen X Li X 2017 Appl. Surf. Sci. 391 72
[8] Zhang S Hang N T Zhang Z Yue H Yang W 2017 Nanomater. (Basel) 7 1
[9] Li R Ren Y Zhao P Wang J Liu J Zhang Y 2019 J. Hazard Mater 365 606
[10] Wu M Gong Y Nie T et al. 2019 J. Mater. Chem. 7 5324
[11] Dong F Zhao Z Xiong T et al. 2013 ACS Appl. Mater Interfaces 5 11392
[12] Xu J Li Y Peng S Lu G Li S 2013 Phys. Chem. Chem. Phys. 15 7657
[13] Zhang L Jing D She X et al. 2014 J. Mater. Chem. 2 2071
[14] Zhang Y Pan Q Chai G et al. 2013 Sci. Rep. 3 1943
[15] Fan X Xing Z Shu Z Zhang L Wang L Shi J 2015 RSC Adv. 5 8323
[16] Xu H Y Wu L C Zhao H Jin L G Qi S Y 2015 PLoS One 10 e0142616
[17] Xin G Meng Y 2013 J. Chem. 2013 1
[18] Ming L Yue H Xu L Chen F 2014 J. Mater. Chem. 2 19145
[19] Yu X Yang X Li G 2018 J. Electron. Mater. 47 672
[20] Mai Oanh L T Hang L T Lai N D et al. 2018 Phys. B: Condens. Matter 532 48
[21] Dai H Gao X Liu E et al. 2013 Diamond Relat. Mater. 38 109
[22] Rehman Iu Movasaghi Z Rehman S 2012 Vibrational Spectroscopy for Tissue Analysis Florida CRC Press 285
[23] Kesarla M K Fuentez-Torres M O Alcudia-Ramos M A et al. 2019 J. Mater. Res. Technol. 8 1628
[24] Shen L Xing Z Zou J et al. 2017 Sci. Rep. 7 41978
[25] Yuan Y Zhang L Xing J et al. 2015 Nanoscale 7 12343